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 12 毫秒
1.
The methyl-accepting chemotaxis proteins (MCPs) are concentrated at the cell poles in an evolutionarily diverse panel of bacteria and an archeon. In elongated cells, the MCPs are located both at the poles and at regions along the length of the cells. Together, these results suggest that MCP location is evolutionarily conserved.  相似文献   

2.
The abundance and diversity of archaeal ammonia monooxygenase subunit A (amoA) genes from hydrothermal vent chimneys at the Juan de Fuca Ridge were investigated. The majority of the retrieved archaeal amoA sequences exhibited identities of less than 95% to those in the GenBank database. Novel ammonia-oxidizing archaea may exist in the hydrothermal vent environments.Ammonia-oxidizing archaea (AOA) may play important roles in carbon and nitrogen cycles in various temperate environments (5, 7, 10, 12, 16). The frequent detection (23, 24) and successful enrichment (2, 6) of thermophilic AOA from terrestrial hot springs suggested a wide distribution of thermophilic AOA in geothermal environments. High concentrations of NH4+ (1, 9, 11) and high rates of ammonia oxidation (9, 22) have been observed at the Juan de Fuca Ridge. However, the presence of AOA in this deep-sea hydrothermal system has not been reported. Here, the abundance and diversity of AOA in three hydrothermal vent chimneys in the Endeavor segment of the Juan de Fuca Ridge were investigated by targeting the conserved amoA genes. This is also the first report on AOA from deep-sea hydrothermal vent chimneys.These vent chimneys were sulfide structures obtained in the fall of 2005 using the submersible Alvin on board the research vessel Atlantis (dive numbers 4143, 4136, and 4148). Chimney 4148 was an active black smoker venting at around 310°C in the Main Endeavor field (47°56.876′N, 129°5.915′W; depth, 2,192 m). Chimney 4143-1 was an active black smoker venting at 316°C in the Mothra field (47°55.424′N, 129°6.533′W; depth, 2,267 m). The outer layers (samples 4148-1A and 4143-1A) of these chimneys were used in this study. The sample from chimney 4136-1 was from a diffusive field (Clambed field) (47°57.909′N, 129°5.443′W; depth, 2,200 m), where the in situ temperature was measured as 29.2°C. The chimney samples were stored at −20°C on board, transported to the home laboratory on dry ice, and stored at −80°C until analyses were performed.Chimney samples were frozen in liquid nitrogen and milled upon thawing. This procedure was repeated three times to break down the solid sample into small particles, which were then mixed with DNA extraction buffer for DNA isolation as described before (25). The obtained crude DNA was purified by an E-Z N.A. Cycle-Pure kit (Omega Bio-Tek Inc., Norcross, GA). PCR amplifications for the archaeal 16S rRNA gene, the crenarchaeal marine group I (MGI) 16S rRNA gene, the archaeal amoA gene, and the bacterial amoA gene followed procedures previously described (Table (Table1)1) (3, 5, 10, 14). Quantitative PCR (Q-PCR) was performed using a model 7500 real-time system (Applied Biosystems, United Kingdom) and a 20-μl reaction mixture that consisted of 1 μl (1 to 10 ng) of DNA as the template, a 0.15 μM concentration of each primer, and 10 μl of Power SYBR green PCR master mix (Applied Biosystems, United Kingdom) with ROX and SYBR green I. The inserted PCR fragments of clones 4143-1A-71 (from the amoA gene library) and 4136-1-4 (from the archaeal 16S rRNA gene library) were amplified and purified to generate standard DNAs for amoA or archaeal 16S rRNA gene quantification. A serial dilution of standard DNAs was performed to generate calibration curves for sample quantification. A melting curve analysis was performed after amplification, and the cycle threshold was set automatically using system 7500 software, version 1.3.

TABLE 1.

PCR primers and procedures used in this study
Target genePrimerSequence (5′→3′)PCR cycle conditionsReference
Archaeal amoAArch-amoAFSTAATGGTCTGGCTTAGACG5 min at 95°C; 30 cycles consisting of 45 s at 94°C, 1 min at 53°C, and 1 min at 72°C; 15 min at 72°CFrancis et al., 2005 (5)
Arch-amoARGCGGCCATCCATCTGTATGT
Archaeal 16S rRNA21FTTCCGGTTGATCCYGCCRG5 min at 95°C; 30 cycles consisting of 30 s at 94°C, 1 min at 54°C, and 1 min at 72°C; 10 min at 72°CDeLong, 1992 (3)
958RYCCGGCGTTGAMTCCAATT
Archaeal 16S rRNA (for Q-PCR)344FACGGGGCGCAGCAGGCGCGA10 min at 50°C, 2 min at 95°C; 40 cycles consisting of 15 s at 95°C and 1 min at 60°C; 15 s at 95°C, 1 min at 60°C, and 15 s at 95°C to make the melting curveØvreås et al., 1998 (15)
518RATTACCGCGGCTGCTGG
Archaeal amoA (for Q-PCR)amo196FGGWGTKCCRGGRACWGCMAC10 min at 50°C and 2 min at 95°C; 40 cycles consisting of 15 s at 95°C and 1 min at 60°C; 15 s at 95°C, 1 min at 60°C, and 15 s at 95°C to make the melting curveTreusch et al., 2005 (20)
amo277RCRATGAAGTCRTAHGGRTADCC
Bacterial amoAAmoA-1FGGGGTTTCTACTGGTGGT5 min at 95°C; 30 cycles consisting of 30 s at 94°C, 45 s at 54°C-50°C, and 45 s at 72°C; 10 min at 72°CStephen et al., 1998 (19)
AmoA-2RCCCCTCKGSAAAGCCTTCTTCRotthauwe et al., 1997 (17)
Crenarchaeal marine group I 16S rRNA771FACGGTGAGGGATGAAAGCT5 min at 95°C; 30 cycles consisting of 30 s at 95°C, 30 s at 54°C, and 30 s at 72°COchsenreiter et al., 2003 (14)
957RCGGCGTTGACTCCAATTG
Open in a separate windowTriplicate PCR products were pooled and clone libraries constructed following the manufacturer''s instructions (Takara Inc., Dalian, China). PCR clones from the libraries were randomly selected for sequencing (Sangon Inc., China). Phylogenetic trees were generated using the PHYLIP package (4) and the maximum-likelihood, neighbor-joining, and maximum-parsimony methods. Bootstrap analysis was used to estimate the reliability of phylogenetic tree constructions (200 replicates). Trees were created using the program Treeview (version 1.6.6).Positive and specific PCR bands were obtained for the archaeal amoA genes from all the three samples, while no PCR band was obtained for the bacterial amoA gene (for the primers and procedures used, see Table Table1).1). In addition, sample 4136-1 was found by Q-PCR analysis to contain the highest number of archaeal amoA genes (with 7.36 ± 0.37 × 104 copies per g of chimney), followed by samples 4143-1A (with 1.88 ± 0.08 × 104 copies per g of chimney) and 4148-1A (with 1.37 ± 0.07 × 102 copies per g of chimney).Clone libraries of archaeal amoA from the three samples were constructed. A total of 93 clones (33 from sample 4136-1, 30 from sample 4143-1A, and 30 from sample 4148-1A) were sequenced and divided into 33 operational taxonomic units (OTUs) based on 99% nucleotide identity. The majority (81.7%) of the retrieved archaeal amoA OTU sequences exhibited relatively low identity (≤94.56%) to other archaeal amoA sequences deposited in GenBank. The phylogenetic relationships among the retrieved amoA and some published amoA sequences are shown in Fig. Fig.1.1. The chimney archaeal amoA sequences fell into five clusters (chimney group I, chimney group II, sediment A-1, and water column A and B clusters), except the sequence of clone 4143-1A-10, which did not fall into any cluster and exhibited the highest identity (90%) to the sequence of clone HB_B_0805A06, which was derived from coastal sediment (18). Chimney group I contained 52 sequences (30 from sample 4148-1A, 11 from sample 4143-1A, and 11 from sample 4136-1); chimney group II contained 23 sequences (20 from sample 4136-1 and 3 from sample 4143-1A). Fourteen sequences from sample 4143-1A grouped into water column A and B clusters (5); and one sequence from sample 4143-1A grouped into the sediment A-1 cluster (13). The sequences from chimney group I exhibited the highest identity (94%) to clone CR-G3N006, derived from a cold seep of the Japan Sea (13). Sequences in chimney group II exhibited the highest identity to clone OA-MA-122 from a water column of a coastal aquarium biofilter, with 84% nucleotide identity (21). The sequences of chimney group II did not cluster with any other sequences. Although showing low bootstrap values (<50%), the chimney group II sequences always clustered into a separate group (Fig. (Fig.1)1) according to different calculation methods, including the maximum-likelihood, neighbor-joining, and maximum-parsimony methods.Open in a separate windowFIG. 1.Phylogenetic tree showing the affiliations of archaeal amoA gene sequences from chimneys (in bold), sediments, soil, water, and the isolated AOA. Bootstrap values were calculated from 200 replications with 585 characters. Maximum-likelihood (left), distance (middle), and parsimony (right) bootstrap values providing ≥50% support are indicated. The bar represents 100 expected substitutions for the archaeal amoA region analyzed. Bacterial amoA sequences were set as the outgroup.Sample 4136-1 contained the highest number of archaeal amoA gene copies. Q-PCR using primers 344F and 518R (15) showed that sample 4136-1 contained 1.10 ± 0.05 × 106 copies of archaeal 16S rRNA genes per g of chimney. Assuming that each crenarchaeal cell possessed only one copy of the amoA gene (8), the AOA constituted at least 6.1% of the archaeal community in sample 4136-1. To explore the potential sources of these amoA sequences in sample 4136-1, an archaeal 16S rRNA clone library was constructed and a total of 82 clones were sequenced. These sequences divided into 20 OTUs based on 98% nucleotide identity. Fifteen OTUs (accounting 76.8% of the total sequences) belonged to hyperthermophilic Desulfurococcales species, and two OTUs (accounting for 15.9% of the total number of sequences) belonged to hyperthermophilic Thermoproteales species of the Crenarchaeota phylum, whereas three OTUs (accounting 7.32% of the total number of sequences) belonged to Thermococcales species of the Euryarchaeota kingdom (Fig. (Fig.2).2). Members of the crenarchaeal MGI, which was thought to be the source of nonthermophilic AOA (6, 8), were not detected in this library. Therefore, PCR using MGI-specific primers was performed to further detect MGI species (for PCR primers and conditions, see Table Table11 and reference 14). MGI species were easily detected in sample 4143-1A, but not in samples 4136-1 and 4148-1A, by direct PCR amplification. A nested PCR method employing generic archaeal 16S rRNA gene primers was then performed for the first round of PCR followed by MGI-selective PCR primers for the second round of PCR. This procedure created a PCR band of the correct size for MGI species from sample 4136-1; that band was later shown by cloning and sequencing to represent an MGI 16S rRNA gene fragment (see Fig. S1 in the supplemental material). The data implied that some of the amoA genes detected in the chimney samples may have come from MGI species; however, to determine the origin of the amoA genes, especially those in the chimney groups I and II, isolation or enrichment of the organisms would be necessary.Open in a separate windowFIG. 2.Phylogenetic tree showing the affiliations of 16S rRNA gene sequences retrieved from hydrothermal vent chimney 4136-1 (in boldface) with selected reference sequences of the Archaea domain. Bootstrap values were calculated from 200 replications with 790 characters. Maximum-likelihood (left), distance (middle), and parsimony (right) bootstrap values providing ≥50% support are indicated. The bar represents 100 expected substitutions for the archaeal 16S rRNA gene analyzed. Bacterial 16S rRNA sequences were set as the outgroup. HWCGIII, hot water crenarchaeotic group III.  相似文献   

3.
4.
Recent studies have highlighted the surprising richness of soil bacterial communities; however, bacteria are not the only microorganisms found in soil. To our knowledge, no study has compared the diversities of the four major microbial taxa, i.e., bacteria, archaea, fungi, and viruses, from an individual soil sample. We used metagenomic and small-subunit RNA-based sequence analysis techniques to compare the estimated richness and evenness of these groups in prairie, desert, and rainforest soils. By grouping sequences at the 97% sequence similarity level (an operational taxonomic unit [OTU]), we found that the archaeal and fungal communities were consistently less even than the bacterial communities. Although total richness levels are difficult to estimate with a high degree of certainty, the estimated number of unique archaeal or fungal OTUs appears to rival or exceed the number of unique bacterial OTUs in each of the collected soils. In this first study to comprehensively survey viral communities using a metagenomic approach, we found that soil viruses are taxonomically diverse and distinct from the communities of viruses found in other environments that have been surveyed using a similar approach. Within each of the four microbial groups, we observed minimal taxonomic overlap between sites, suggesting that soil archaea, bacteria, fungi, and viruses are globally as well as locally diverse.  相似文献   

5.
We have sequenced the genome and identified the structural proteins and lipids of the novel membrane-containing, icosahedral virus P23-77 of Thermus thermophilus. P23-77 has an ∼17-kb circular double-stranded DNA genome, which was annotated to contain 37 putative genes. Virions were subjected to dissociation analysis, and five protein species were shown to associate with the internal viral membrane, while three were constituents of the protein capsid. Analysis of the bacteriophage genome revealed it to be evolutionarily related to another Thermus phage (IN93), archaeal Halobacterium plasmid (pHH205), a genetic element integrated into Haloarcula genome (designated here as IHP for integrated Haloarcula provirus), and the Haloarcula virus SH1. These genetic elements share two major capsid proteins and a putative packaging ATPase. The ATPase is similar with the ATPases found in the PRD1-type viruses, thus providing an evolutionary link to these viruses and furthering our knowledge on the origin of viruses.Three-dimensional structures of the major capsid proteins, as well as the architecture of the virion and the sequence similarity of putative genome packaging ATPases, have revealed unexpected evolutionary connection between virus families. Viruses infecting hosts residing in different domains of life (Bacteria, Archaea, and Eukarya) share common structural elements and possibly also ways to package the viral genome (8, 13, 41). It has been proposed that the set of genes responsible for virion assembly is a hallmark of the virus and is designated as the innate viral “self,” which may retain its identity through evolutionary times (5). Based on this, it is proposed that viruses can be classified into lineages that span the different domains of life. Therefore, the studies of new virus isolates might provide insights into the events that led to the origin of viruses and maybe even the origin of life itself (34, 40). However, viruses are known to be genetic mosaics (28), and these structural lineages therefore do not reflect the evolutionary history of all genes in a given virus. For example, the genome replication strategies vary significantly even in the currently established lineages (41) and, consequently, a structural approach does not point out to a specific form of replication in the ancestor. Nevertheless, as the proposal for a viral self is driven from information on viral structures and pathways of genome encapsidation, the ancestral form of the self was likely to be composed of a protective coat and the necessary mechanisms to incorporate the genetic material within the coat.Viruses structurally related to bacteriophage PRD1, a phage infecting gram-negative bacteria, have been identified in all three domains of life, and the lineage hypothesis was first proposed based on structural information on such viruses. Initially, PRD1 and human adenovirus were proposed to originate from a common ancestor mainly due to the same capsid organization (T=25) and the major coat protein topology, the trimeric double β-barrel fold (12). In addition, these viruses share a common vertex organization and replication mechanism (20, 31, 53, 63). PRD1 is an icosahedral virus with an inner membrane, whereas adenovirus lacks the membrane. Later, many viruses with similar double β-barrel fold in the major coat protein have been discovered and included to this viral lineage. For example, the fold is present in Paramecium bursaria Chlorella virus 1 (56) of algae, Bam35 (45) of gram-positive bacteria, PM2 (2) of gram-negative marine bacteria, and Sulfolobus turreted icosahedral virus (STIV) (38) of an archaeal host. Moreover, genomic analyses have revealed a common set of genes in a number of nucleocytoplasmic large DNA viruses. Chilo iridescent virus and African swine fever virus 1 are related to Paramecium bursaria Chlorella virus 1 and most probably share structural similarity to PRD1-type viruses (13, 30, 31, 68). The largest known viruses, represented by mimivirus and poxvirus, may also belong to this lineage (29, 77). Two euryarchaeal proviruses, TKV4 and MVV, are also proposed to belong to this lineage based on bioinformatic searches (42). The proposed PRD1-related viruses share the same basic architectural principles despite major differences in the host organisms and particle and genome sizes (1, 2, 38, 56). PM2, for example, has a genome of only 10 kbp, whereas mimivirus (infecting Acanthamoeba polyphaga) double-stranded DNA (dsDNA) genome is 1.2 Mbp in size (59).How many virion structure-based lineages might there be? This obviously relates to the number of protein folds that have the properties needed to make viral capsids. It has been noted that, in addition to PRD1-type viruses, at least tailed bacterial and archaeal viruses, as well as herpesviruses, share the same coat protein fold. Also, certain dsRNA viruses seem to have structural and functional similarities, although their hosts include bacteria and yeasts, as well as plants and animals (6, 18, 19, 27, 55, 60, 74). Obviously, many structural principles to build a virus capsid exist, and it has been suggested that especially geothermally heated environments have preserved many of the anciently formed virus morphotypes (35).Thermophilic dsDNA bacteriophage P23-77 was isolated from an alkaline hot spring in New Zealand on Thermus thermophilus (17) ATCC 33923 (deposited as Thermus flavus). P23-77 was shown to have an icosahedral capsid and possibly an internal membrane but no tail (81). Previously, another Thermus virus, IN93, with a similar morphology has been described (50). IN93 was inducible from a lysogenic strain of Thermus aquaticus TZ2, which was isolated from hot spring soil in Japan. Recently, P23-77 was characterized in more detail (33). It has an icosahedral protein coat, organized in a T=28 capsid lattice (21). The presence of an internal membrane was confirmed, and lipids were shown to be constituents of the virion. Ten structural proteins were identified, with apparent molecular masses ranging from 8 to 35 kDa. Two major protein species with molecular masses of 20 and 35 kDa were proposed to make the capsomers, one forming the hexagonal building blocks and the other the two towers that decorate the capsomer bases (33). Surprisingly, P23-77 is structurally closest to the haloarchaeal virus SH1, which is the only other example of a T=28 virion architecture (32, 33). In both cases it was proposed that the capsomers are made of six single β-barrels opposing the situation with the other structurally related viruses where the hexagonal capsomers are made of three double β-barrel coat protein monomers (8).In the present study we analyze the dsDNA genome of P23-77. Viral membrane proteins and those associated with the capsid were identified by virion dissociation studies. The protein chemistry data and genome annotation are consistent with the results of the disruption studies. A detailed analysis of the lipid composition of P23-77 and its T. thermophilus host was carried out. The data collected here reveal additional challenges in attempts to generate viral lineages based on the structural and architectural properties of the virion.  相似文献   

6.
Alvinella pompejana is a polychaetous annelid that inhabits high-temperature environments associated with active deep-sea hydrothermal vents along the East Pacific Rise. A unique and diverse epibiotic microflora with a prominent filamentous morphotype is found associated with the worm's dorsal integument. A previous study established the taxonomic positions of two epsilon proteobacterial phylotypes, 13B and 5A, which dominated a clone library of 16S rRNA genes amplified by PCR from the epibiotic microbial community of an A. pompejana specimen. In the present study deoxyoligonucleotide PCR primers specific for phylotypes 13B and 5A were used to demonstrate that these phylotypes are regular features of the bacterial community associated with A. pompejana. Assaying of other surfaces around colonies of A. pompejana revealed that phylotypes 13B and 5A are not restricted to A. pompejana. Phylotype 13B occurs on the exterior surfaces of other invertebrate genera and rock surfaces, and phylotype 5A occurs on a congener, Alvinella caudata. The 13B and 5A phylotypes were identified and localized on A. pompejana by in situ hybridization, demonstrating that these two phylotypes are, in fact, the prominent filamentous bacteria on the dorsal integument of A. pompejana. These findings indicate that the filamentous bacterial symbionts of A. pompejana are epsilon Proteobacteria which do not have an obligate requirement for A. pompejana.  相似文献   

7.
In the present study, the influence of the land use intensity on the diversity of ammonia oxidizing bacteria (AOB) and archaea (AOA) in soils from different grassland ecosystems has been investigated in spring and summer of the season (April and July). Diversity of AOA and AOB was studied by TRFLP fingerprinting of amoA amplicons. The diversity from AOB was low and dominated by a peak that could be assigned to Nitrosospira. The obtained profiles for AOB were very stable and neither influenced by the land use intensity nor by the time point of sampling. In contrast, the obtained patterns for AOA were more complex although one peak that could be assigned to Nitrosopumilus was dominating all profiles independent from the land use intensity and the sampling time point. Overall, the AOA profiles were much more dynamic than those of AOB and responded clearly to the land use intensity. An influence of the sampling time point was again not visible. Whereas AOB profiles were clearly linked to potential nitrification rates in soil, major TRFs from AOA were negatively correlated to DOC and ammonium availability and not related to potential nitrification rates.  相似文献   

8.
The study of of the distribution of microorganisms through space (and time) allows evaluation of biogeographic patterns, like the species-area index (z). Due to their high dispersal ability, high reproduction rates and low rates of extinction microorganisms tend to be widely distributed, and they are thought to be virtually cosmopolitan and selected primarily by environmental factors. Recent studies have shown that, despite these characteristics, microorganisms may behave like larger organisms and exhibit geographical distribution. In this study, we searched patterns of spatial diversity distribution of bacteria and archaea in a contiguous environment. We collected 26 samples of a lake sediment, distributed in a nested grid, with distances between samples ranging from 0.01 m to 1000 m. The samples were analyzed using T-RFLP (Terminal restriction fragment length polymorphism) targeting mcrA (coding for a subunit of methyl-coenzyme M reductase) and the genes of Archaeal and Bacterial 16S rRNA. From the qualitative and quantitative results (relative abundance of operational taxonomic units) we calculated the similarity index for each pair to evaluate the taxa-area and distance decay relationship slopes by linear regression. All results were significant, with mcrA genes showing the highest slope, followed by Archaeal and Bacterial 16S rRNA genes. We showed that the microorganisms of a methanogenic community, that is active in a contiguous environment, display spatial distribution and a taxa-area relationship.  相似文献   

9.
This paper reports the discovery of anaerobic respiration on tellurate by bacteria isolated from deep ocean (1,543 to 1,791 m) hydrothermal vent worms. The first evidence for selenite- and vanadate-respiring bacteria from deep ocean hydrothermal vents is also presented. Enumeration of the anaerobic metal(loid)-resistant microbial community associated with hydrothermal vent animals indicates that a greater proportion of the bacterial community associated with certain vent fauna resists and reduces metal(loid)s anaerobically than aerobically, suggesting that anaerobic metal(loid) respiration might be an important process in bacteria that are symbiotic with vent fauna. Isolates from Axial Volcano and Explorer Ridge were tested for their ability to reduce tellurate, selenite, metavanadate, or orthovanadate in the absence of alternate electron acceptors. In the presence of metal(loid)s, strains showed an ability to grow and produce ATP, whereas in the absence of metal(loid)s, no growth or ATP production was observed. The protonophore carbonyl cyanide m-chlorophenylhydrazone depressed metal(loid) reduction. Anaerobic tellurate respiration will be a significant component in describing biogeochemical cycling of Te at hydrothermal vents.  相似文献   

10.
11.
We review and update the work on genetic elements, e.g., viruses and plasmids (excluding IS elements and transposons) in the kingdom Crenarchaeota (Thermoproteales and Sulfolobales) and the orders Thermococcales and Thermoplasmales in the kingdom Euryachaeota of the archael domain, including unpublished data from our laboratory. The viruses of Crenarchaeota represent four novel virus families. The Fuselloviridae represented by SSV1 of S. shibatae and relatives in other Sulfolobus strains have the form of a failed spindle. The envelope is highly hydrophobic. The DNA is double-stranded and circular. Members of this group have also been found in Methanococcus and Haloarcula. The Lipothrixviridae (e.g., T TV1 to 3) have the form of flexible filaments. They have a core containing linear double-stranded DNA and DNA-binding proteins which is wrapped into a lipid membrane. The ‘Bacilloviridae’ (e.g., TTV4 and SIRV) are stiff rods lacking this membrane, but also featuring linear double-stranded DNA and DNA-binding proteins. Both virus type carry on both ends structures involved in the attachment to receptors. Both types are represented in Thermoproteus and Sulfolobus. The droplet-formed novel Sulfolobus virus SNDV represents the ‘Guttaviridae’ containing circular double-stranded DNA. Though head and tail viruses distantly resembling T phages or lambdoid phages were seen electronmicroscopically in solfataric water samples, no such virus has so far been isolated. SSV1 is temperate, TTV1 causes lysis after induction, the other viruses found so far exist in carrier states. The hosts of all but TTV1 survive virus production. We discuss the implications of the nature of these viruses for understanding virus evolution. The plasmids found so far range in size from 4.5 kb to about 40 kb. Most of them occur in high copy number, probably due to the way of their detection. Most are cryptic, pNOB8 is conjugative, the widespread pDL10 alleviates in an unknown way autotrophic growth of its host Desulfurolobus by sulfur reduction. The plasmid pTIK4 appears to encode a killer function. pNOB8 has been used as a vector for the transfer of the lac S (β-galactosidase) gene into a mutant of S. solfataricus.  相似文献   

12.
Reconstruction of the Tree of Life is a central goal in biology. Although numerous novel phyla of bacteria and archaea have recently been discovered, inconsistent phylogenetic relationships are routinely reported, and many inter-phylum and inter-domain evolutionary relationships remain unclear. Here, we benchmark different marker genes often used in constructing multidomain phylogenetic trees of bacteria and archaea and present a set of marker genes that perform best for multidomain trees constructed from concatenated alignments. We use recently-developed Tree Certainty metrics to assess the confidence of our results and to obviate the complications of traditional bootstrap-based metrics. Given the vastly disparate number of genomes available for different phyla of bacteria and archaea, we also assessed the impact of taxon sampling on multidomain tree construction. Our results demonstrate that biases between the representation of different taxonomic groups can dramatically impact the topology of resulting trees. Inspection of our highest-quality tree supports the division of most bacteria into Terrabacteria and Gracilicutes, with Thermatogota and Synergistota branching earlier from these superphyla. This tree also supports the inclusion of the Patescibacteria within the Terrabacteria as a sister group to the Chloroflexota instead of as a basal-branching lineage. For the Archaea, our tree supports three monophyletic lineages (DPANN, Euryarchaeota, and TACK/Asgard), although we note the basal placement of the DPANN may still represent an artifact caused by biased sequence composition. Our findings provide a robust and standardized framework for multidomain phylogenetic reconstruction that can be used to evaluate inter-phylum relationships and assess uncertainty in conflicting topologies of the Tree of Life.  相似文献   

13.
In the past two years, archaeal genomics has achieved several breakthroughs. On the evolutionary front the most exciting development was the sequencing and analysis of the genome of Nanoarchaeum equitans, a tiny parasitic organism that has only approximately 540 genes. The genome of Nanoarchaeum shows signs of extreme rearrangement including the virtual absence of conserved operons and the presence of several split genes. Nanoarchaeum is distantly related to other archaea, and it has been proposed to represent a deep archaeal branch that is distinct from Euryarchaeota and Crenarchaeota. This would imply that many features of its gene repertoire and genome organization might be ancestral. However, additional genome analysis has provided a more conservative suggestion - that Nanoarchaeum is a highly derived euryarchaeon. Also there have been substantial developments in functional genomics, including the discovery of the elusive aminoacyl-tRNA synthetase that is involved in both the biosynthesis of cysteine and its incorporation into proteins in methanogens, and the first experimental validation of the predicted archaeal exosome.  相似文献   

14.
The accurate copying of genetic information in the double helix of DNA is essential for inheritance of traits that define the phenotype of cells and the organism. The core machineries that copy DNA are conserved in all three domains of life: bacteria, archaea, and eukaryotes. This article outlines the general nature of the DNA replication machinery, but also points out important and key differences. The most complex organisms, eukaryotes, have to coordinate the initiation of DNA replication from many origins in each genome and impose regulation that maintains genomic integrity, not only for the sake of each cell, but for the organism as a whole. In addition, DNA replication in eukaryotes needs to be coordinated with inheritance of chromatin, developmental patterning of tissues, and cell division to ensure that the genome replicates once per cell division cycle.The genetic information within the cells of our body is stored in the double helix of DNA, a long cylinderlike structure with a radius that is only 10 Å or one billionth of a meter but can be of considerable length. A single DNA molecule within a bacterium that grows in our gut flora is approximately 5 million base pairs in length and when stretched out, is about 1.6 mm in length, roughly the diameter of a pinhead. In contrast, the single DNA molecule in the largest human chromosome is 245,203,898 base pairs or about 8.33 cm long. The entire human genome, consisting of its 24 different chromosomes in a male is about 3 billion base pairs or 1 m long. Each cell in our body, with rare exceptions, contains two copies of the genome and thus 2 m of total DNA. Thus the scale and complexity of duplicating genomes is remarkable. For example, ∼2200 human cells can sit on the top of a 1.5 mm pinhead and when extracted and laid out in a line, the DNA from these cells would be ∼4.5 km (2.8 miles) long. In our body, about 500–700 million new blood cells are born every minute in the bone marrow (Doulatov et al. 2012), containing a total of about 1 million km of DNA, or enough DNA to wrap around the equator of the earth 25 times. Thus DNA replication is a serious business in our body, occurring from the time that a fertilized egg first begins duplicating DNA to yield the many trillions of cells that make up an adult body and continuing in all tissues of the adult body throughout our life. The amount of DNA duplicated in an entire human body represents an unimaginable amount of information transfer. Moreover, each round of duplication needs to be highly accurate, making one mistake in less than 100 million bases copied per cell division. How copying of the double helix occurs and how it is so highly accurate is the topic of this collection. Inevitably the processes of accurate copying of the genome can go awry, yielding mutations that affect our lives, and thus the collection outlines the disorders that accelerate human disease.However, the problem of copying DNA is much more complicated than indicated above. The 2 m of DNA in each human cell is wrapped up with histone proteins within the cell’s nucleus that is only about 5 μm wide, presenting a compaction in DNA length of about 2 million-fold. How can the copying process deal with the fact that the DNA is wrapped around proteins and scrunched into a volume that creates a spatial organization problem of enormous magnitude? Not only is the DNA copied, but the proteins associated with the DNA need to be duplicated, along with all the chemical modifications attached to DNA and histones that greatly influence developmental patterning of gene expression. The protein machineries that replicate DNA and duplicate proteins within the chromosomes are some of the most complex and intriguing machineries known. Furthermore, the regulations of the processes are some of the most complex because they need to ensure that each DNA molecule in each chromosome is copied once, and only once each time before a cell divides. Errors in the regulation of DNA replication lead to accelerated mutation rates, often associated with increased rates of cancer and other diseases.The process of accurately copying a genome can be broken down into various subprocesses that combine to provide efficient genome duplication. Central to the entire process is the machinery that actually copies the DNA with high fidelity, including proteins that start the entire process and the proteins that actually copy one helix to produce two. Superimposed on this fundamental process are mechanisms that detect and repair errors and damage to the DNA. Also associated with the DNA replication apparatus are the proteins that ensure that the histone proteins and their modifications in chromatin are inherited along with the DNA. Finally, other machineries cooperate with the DNA replication apparatus to ensure that the resulting two DNA molecules, the sister chromatids, are tethered together until the cell completes duplicating all of its DNA and segregates the sister chromatids evenly to the two daughter cells. Only by combining all of these processes can genetic inheritance ensure that each cell has a faithful copy of its parent’s genome.  相似文献   

15.
Specimens of alvinellid polychaetes and their tubes were collected in the Parigo hydrothermal vent field on the East Pacific Rise (13°N) in October and November 1987. Heterotrophic bacterial strains were isolated on metal-amended media from the tube and dorsal integument of one specimen of Alvinella pompejana, from the dorsal integument of another from the whole integument of a specimen of Alvinella caudata, and from undetermined alvinellid tubes. The strains were characterized and tested for susceptibility to five heavy metals by using a microdilution method for MIC determinations. All strains were gram-negative rods. Most of them were characterized by the ability to degrade Tween 80 and gelatin and to produce hydrogen sulfide from cysteine. Numerous strains, from all sample origins, displayed resistance to cadmium, zinc, arsenate, and silver and tolerated high amounts of copper. Metal resistance was exhibited by 92.3% of the total isolates. The occurrence of multiply resistant bacteria may demonstrate an adaptation of alvinellid-associated microflora to the general enrichment of metals in the hydrothermal vent environment.  相似文献   

16.
Comparative biochemistry of Archaea and Bacteria.   总被引:11,自引:0,他引:11  
This review compares exemplary molecular and metabolic features of Archaea and Bacteria in terms of phylogenetic aspects. The results of the comparison confirm the coherence of the Archaea as postulated by Woese. Archaea and Bacteria share many basic features of their genetic machinery and their central metabolism. Similarities and distinctions allow projections regarding the nature of the common ancestor and the process of lineage diversification.  相似文献   

17.
Microorganisms are responsible for multiple antibiotic resistances that have been associated with resistance/tolerance to heavy metals, with consequences to public health. Many genes conferring these resistances are located on mobile genetic elements, easily exchanged among phylogenetically distant bacteria. The objective of the present work was to isolate arsenic-, antimonite-, and antibiotic-resistant strains and to determine the existence of plasmids harboring antibiotic/arsenic/antimonite resistance traits in phenotypically resistant strains, in a nonanthropogenically impacted environment. The hydrothermal Lucky Strike field in the Azores archipelago (North Atlantic, between 11°N and 38°N), at the Mid-Atlantic Ridge, protected under the OSPAR Convention, was sampled as a metal-rich pristine environment. A total of 35 strains from 8 different species were isolated in the presence of arsenate, arsenite, and antimonite. ACR3 and arsB genes were amplified from the sediment''s total DNA, and 4 isolates also carried ACR3 genes. Phenotypic multiple resistances were found in all strains, and 7 strains had recoverable plasmids. Purified plasmids were sequenced by Illumina and assembled by EDENA V3, and contig annotation was performed using the “Rapid Annotation using the Subsystems Technology” server. Determinants of resistance to copper, zinc, cadmium, cobalt, and chromium as well as to the antibiotics β-lactams and fluoroquinolones were found in the 3 sequenced plasmids. Genes coding for heavy metal resistance and antibiotic resistance in the same mobile element were found, suggesting the possibility of horizontal gene transfer and distribution of theses resistances in the bacterial population.  相似文献   

18.
The complete genome sequence of the thermophilic sulphur-reducing bacterium, Deferribacter desulfuricans SMM1, isolated from a hydrothermal vent chimney has been determined. The genome comprises a single circular chromosome of 2 234 389 bp and a megaplasmid of 308 544 bp. Many genes encoded in the genome are most similar to the genes of sulphur- or sulphate-reducing bacterial species within Deltaproteobacteria. The reconstructed central metabolisms showed a heterotrophic lifestyle primarily driven by C1 to C3 organics, e.g. formate, acetate, and pyruvate, and also suggested that the inability of autotrophy via a reductive tricarboxylic acid cycle may be due to the lack of ATP-dependent citrate lyase. In addition, the genome encodes numerous genes for chemoreceptors, chemotaxis-like systems, and signal transduction machineries. These signalling networks may be linked to this bacterium''s versatile energy metabolisms and may provide ecophysiological advantages for D. desulfuricans SSM1 thriving in the physically and chemically fluctuating environments near hydrothermal vents. This is the first genome sequence from the phylum Deferribacteres.  相似文献   

19.
Comparative genomics has revealed that variations in bacterial and archaeal genome DNA sequences cannot be explained by only neutral mutations. Virus resistance and plasmid distribution systems have resulted in changes in bacterial and archaeal genome sequences during evolution. The restriction-modification system, a virus resistance system, leads to avoidance of palindromic DNA sequences in genomes. Clustered, regularly interspaced, short palindromic repeats (CRISPRs) found in genomes represent yet another virus resistance system. Comparative genomics has shown that bacteria and archaea have failed to gain any DNA with GC content higher than the GC content of their chromosomes. Thus, horizontally transferred DNA regions have lower GC content than the host chromosomal DNA does. Some nucleoid-associated proteins bind DNA regions with low GC content and inhibit the expression of genes contained in those regions. This form of gene repression is another type of virus resistance system. On the other hand, bacteria and archaea have used plasmids to gain additional genes. Virus resistance systems influence plasmid distribution. Interestingly, the restriction-modification system and nucleoid-associated protein genes have been distributed via plasmids. Thus, GC content and genomic signatures do not reflect bacterial and archaeal evolutionary relationships.  相似文献   

20.
Of the many pathogens that infect humans and animals, a large number use cells of the host organism as protected sites for replication. To reach the relevant intracellular compartments, they take advantage of the endocytosis machinery and exploit the network of endocytic organelles for penetration into the cytosol or as sites of replication. In this review, we discuss the endocytic entry processes used by viruses and bacteria and compare the strategies used by these dissimilar classes of pathogens.Many of the most widespread and devastating diseases in humans and livestock are caused by viruses and bacteria that enter cells for replication. Being obligate intracellular parasites, viruses have no choice. They must transport their genome to the cytosol or nucleus of infected cells to multiply and generate progeny. Bacteria and eukaryotic parasites do have other options; most of them can replicate on their own. However, some have evolved to take advantage of the protected environment in the cytosol or in cytoplasmic vacuoles of animal cells as a niche favorable for growth and multiplication. In both cases (viruses and intracellular bacteria), the outcome is often destructive for the host cell and host organism. The mortality and morbidity caused by infectious diseases worldwide provide a strong rationale for research into pathogen–host cell interactions and for pursuing the detailed mechanisms of transmission and dissemination. The study of viruses and bacteria can, moreover, provide invaluable insights into fundamental aspects of cell biology.Here, we focus on the mechanisms by which viral and bacterial pathogens exploit the endocytosis machinery for host cell entry and replication. Among recent reviews on this topic, dedicated uniquely to either mammalian viruses or bacterial pathogens, we recommend the following: Cossart and Sansonetti (2004); Pizarro-Cerda and Cossart (2006); Kumar and Valdivia (2009); Cossart and Roy (2010); Mercer et al. (2010b); Grove and Marsh (2011); Kubo et al. (2012); Vazquez-Calvo et al. (2012a); Sun et al. (2013).The term “endocytosis” is used herein in its widest sense, that is, to cover all processes whereby fluid, solutes, ligands, and components of the plasma membrane as well as particles (including pathogenic agents) are internalized by cells through the invagination of the plasma membrane and the scission of membrane vesicles or vacuoles. This differs from current practice in the bacterial pathogenesis field, where the term “endocytosis” is generally reserved for the internalization of molecules or small objects, whereas the uptake of bacteria into nonprofessional phagocytes is called “internalization” or “bacterial-induced phagocytosis.” In addition, the term “phagocytosis” is reserved for internalization of bacteria by professional phagocytes (macrophages, polymorphonuclear leucocytes, dendritic cells, and amoebae), a process that generally but not always leads to the destruction of the ingested bacteria (Swanson et al. 1999; May and Machesky 2001; Henry et al. 2004; Zhang et al. 2010). With a few exceptions, we will not discuss phagocytosis of bacteria or the endocytosis of protozoan parasites such as Toxoplasma and Plasmodium (Robibaro et al. 2001).  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