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 732 毫秒
1.
Magnetotactic bacteria comprise a phylogenetically diverse group that is capable of synthesizing intracellular magnetic particles. Although various morphotypes of magnetotactic bacteria have been observed in the environment, bacterial strains available in pure culture are currently limited to a few genera due to difficulties in their enrichment and cultivation. In order to obtain genetic information from uncultured magnetotactic bacteria, a genome preparation method that involves magnetic separation of cells, flow cytometry, and multiple displacement amplification (MDA) using φ29 polymerase was used in this study. The conditions for the MDA reaction using samples containing 1 to 100 cells were evaluated using a pure-culture magnetotactic bacterium, “Magnetospirillum magneticum AMB-1,” whose complete genome sequence is available. Uniform gene amplification was confirmed by quantitative PCR (Q-PCR) when 100 cells were used as a template. This method was then applied for genome preparation of uncultured magnetotactic bacteria from complex bacterial communities in an aquatic environment. A sample containing 100 cells of the uncultured magnetotactic coccus was prepared by magnetic cell separation and flow cytometry and used as an MDA template. 16S rRNA sequence analysis of the MDA product from these 100 cells revealed that the amplified genomic DNA was from a single species of magnetotactic bacterium that was phylogenetically affiliated with magnetotactic cocci in the Alphaproteobacteria. The combined use of magnetic separation, flow cytometry, and MDA provides a new strategy to access individual genetic information from magnetotactic bacteria in environmental samples.Magnetotactic bacteria synthesize nanosized intracellular magnetic particles, also referred to as magnetosomes, by accumulating iron ions from the environment. Since the first report on the identification of magnetotactic bacteria (2), the morphological and phylogenetic diversity of these organisms has been observed in various aquatic environments (12, 25, 27, 30). However, bacterial strains available in pure culture are currently limited to a few genera. Desulfovibrio magneticus strain RS-1 is the only isolate of magnetotactic bacteria that is classified among the Deltaproteobacteria (13, 23), while Magnetospirillum spp., marine magnetic vibrio strain MV-1, and “Magnetococcus strain MC-1” are phylogenetically affiliated within the Alphaproteobacteria group (24, 27). This limitation is mainly because not much is known about their metabolic requirements, culturing conditions, and obligate coculture requirements.Isolation and enrichment of magnetotactic bacteria are generally conducted by applying a magnetic field to a container containing a sediment sample from the environment. The capillary racetrack method is a highly selective enrichment technique that separates magnetotactic bacteria from other contaminants (31). The magnetic separation method that involves the use of a large glass apparatus is efficient and suitable for analyzing samples containing more than 100 ml of sediment and water (12, 16). These techniques have been applied to investigate community structure and phylogenetic diversity of uncultured magnetotactic bacteria in the environment based on 16S rRNA analyses (3, 7, 26, 29). In a recent study, DNA isolation enabling gene cloning was examined by magnetically collecting a large number of magnetotactic cells from environmental samples, and two gene fragments, probably containing parts of magnetosome islands (MAIs) derived from magnetotactic bacteria of the Alphaproteobacteria, were identified (12). However, this approach allows only for sequence gene information to be obtained from a heterogeneous bacterial community in the sample.Multiple displacement amplification (MDA) can generate microgram quantities of high-quality DNA sample from a few femtograms of DNA template (5, 6). We previously revealed that MDA is a powerful tool for whole-genome amplification from the metagenome of an uncultured bacterial community (32). Studies have been conducted to determine the efficacy of MDA for analyzing genomic DNA preparations from a limited number of bacterial cells (14, 17, 21, 22, 28). Complete genomic sequencing of an uncultured gut symbiont in termites has been achieved using MDA products amplified from approximately 1,000 cells (9). Partial genome sequencing using MDA products from a single uncultured cell has also been reported (17, 22). Such targeted genome analyses using MDA products from a single cell or genetically identical microorganisms is advantageous because it allows the assignment of individual genes to the corresponding microorganisms.In this study, an improved genome preparation method involving racetrack purification and flow cytometry followed by MDA was investigated by using a small number of uncultured magnetotactic bacteria. This method can be used for the identification of new genes from rare magnetotactic bacteria in environmental samples.  相似文献   

2.
In this report, we describe the selective cloning of large DNA fragments from magnetotactic metagenomes from various aquatic habitats. This was achieved by a two-step magnetic enrichment which allowed the mass collection of environmental magnetotactic bacteria (MTB) virtually free of nonmagnetic contaminants. Four fosmid libraries were constructed and screened by end sequencing and hybridization analysis using heterologous magnetosome gene probes. A total of 14 fosmids were fully sequenced. We identified and characterized two fosmids, most likely originating from two different alphaproteobacterial strains of MTB that contain several putative operons with homology to the magnetosome island (MAI) of cultivated MTB. This is the first evidence that uncultivated MTB exhibit similar yet differing organizations of the MAI, which may account for the diversity in biomineralization and magnetotaxis observed in MTB from various environments.Magnetotactic bacteria (MTB) synthesize magnetosomes, which are membrane-enclosed organelles comprising crystals of magnetite (Fe3O4) or, less commonly, greigite (Fe3S4) (3) that are aligned in intracellular chains along dedicated cytoskeletal structures (26, 36, 38). Magnetic alignment along the magnetic field lines of the earth facilitates navigation in the stratified environment within freshwater and marine sediments (3, 13). MTB do not form a coherent phylogenetic group, but the trait of magnetotaxis is found in species within different phylogenetic clades, including Alphaproteobacteria, Deltaproteobacteria, Gammaproteobacteria, and the Nitrospira phylum (1, 3, 10, 41). Different species produce magnetosome crystals with a multitude of different morphologies displaying a broad variety of intracellular arrangements, including one, two, or multiple chains (3, 14). The perfectly shaped magnetosome crystals and highly ordered chain structures cannot be synthesized by chemical methods as yet. Therefore, an understanding of the genetic mechanisms controlling magnetosome formation is also of great interest for the inorganic production of advanced magnetic nanomaterials (3, 13, 28).Most genes controlling magnetosome formation and magnetotaxis in Magnetospirillum gryphiswaldense and other freshwater magnetospirilla are clustered within four major operons (mamAB, mamGFDC, mms6, and mamXY) (18, 34, 37, 49) that are part of a large genomic magnetosome island (MAI) (49). It was recently shown that the MAI is also conserved in marine MTB, including the MV-1 magnetotactic vibrio strain and the MC-1 magnetic coccus strain. The homologous genomic regions display similar gene contents and, to a lesser extent, a conserved gene synteny (23). It has been suggested that the MAI was transferred horizontally between different MTB (37). However, the divergence between the MAI regions of strain MV-1, strain MC-1, and the magnetospirilla suggests that the events of horizontal gene transfer (HGT) did not occur very recently.Despite continued efforts by many laboratories, the majority of MTB are still not available in pure culture. In particular, the huge diversity of uncultivated species with respect to different morpho- and phylotypes and, in particular, magnetosome crystal shapes is not nearly fully represented by cultivated species. Thus, understanding of the genetic diversity of the magnetotaxis and magnetosome biosynthetic machinery has to rely on culture-independent techniques such as the metagenomic analysis of environmental MTB (24).It has been demonstrated that single genes and even entire operons can be cloned and functionally expressed from uncultivated soil or marine bacteria by using large insert libraries that provide contiguous sections from single organisms (4, 21, 22). The potential to identify and clone genes for metabolic pathways with relevance for biotechnological applications has already been demonstrated in metagenomic projects, such as the identification of polyketide synthase genes from microbial consortia of marine sponges (25) or other environmental samples (8, 31). The cost of sequencing and the challenges that are associated with the management of vast datasets, however, preclude comprehensive genomic studies of highly complex communities. Consequently, approaches that are based on the analysis of a group of bacteria with reduced species diversity are favored. This requires that the sample material is enriched for the target organisms before DNA preparation, for example, by flow sorting, centrifugation, or other physical enrichment techniques (32, 42) or by focusing on natural communities with reduced species diversity (48).Unlike other uncultivated bacteria, MTB exhibit magnetically directed swimming behavior, which enables their selective enrichment from environmental samples without the need of cultivation (16). This approach was utilized in a number of earlier studies uncovering the morphological and phylogenetic diversity of MTB found in environmental populations (10, 43, 45-47). However, these investigations were confined to PCR-based analysis of 16S rRNA genes, ultrastructural studies, and fluorescence in situ hybridization.In this study we used an improved magnetic collection technique to selectively harvest large numbers of uncultivated MTBs, which allowed the extraction of genomic DNA for the construction of large insert metagenomic libraries from different aquatic habitats. Large parts of the MAI from two uncultivated MTB were identified by hybridization using heterologous magnetosome gene probes and end sequencing. We demonstrate for the first time that uncultivated MTB exhibit a clustered organization of magnetosome genes which resembles that of cultivated species and yet displays variations that may account for the observed diversity in biomineralization and magnetotaxis in MTB from various environments. The levels of similarity between and synteny of magnetosome genes of uncultivated and cultivated MTB provide further evidence for HGT.  相似文献   

3.
Magnetosome biomineralization and magnetotaxis in magnetotactic bacteria are controlled by numerous, mostly unknown gene functions that are predominantly encoded by several operons located within the genomic magnetosome island (MAI). Genetic analysis of magnetotactic bacteria has remained difficult and requires the development of novel tools. We established a Cre-lox-based deletion method which allows the excision of large genomic fragments in Magnetospirillum gryphiswaldense. Two conjugative suicide plasmids harboring lox sites that flanked the target region were subsequently inserted into the chromosome by homologous recombination, requiring only one single-crossover event, respectively, and resulting in a double cointegrate. Excision of the targeted chromosomal segment that included the inserted plasmids and their resistance markers was induced by trans expression of Cre recombinase, which leaves behind a scar of only a single loxP site. The Cre helper plasmid was then cured from the deletant strain by relief of antibiotic selection. We have used this method for the deletion of 16.3-kb, 61-kb, and 67.3-kb fragments from the genomic MAI, either in a single round or in subsequent rounds of deletion, covering a region of approximately 87 kb that comprises the mamAB, mms6, and mamGFDC operons. As expected, all mutants were Mag and some were Mot; otherwise, they showed normal growth patterns, which indicates that the deleted region is not essential for viability in the laboratory. The method will facilitate future functional analysis of magnetosome genes and also can be utilized for large-scale genome engineering in magnetotactic bacteria.Magnetosomes are unique membrane-enveloped organelles that are formed by magnetotactic bacteria (MTB) for magnetic navigation (2, 37). The mechanism of magnetosome formation is within the focus of a multidisciplinary interest and has relevance for biotechnological applications (5). It has been recognized that the biomineralization of inorganic magnetite crystals and their assembly into highly ordered magnetosome chains are under strict genetic control. Recent studies combining proteomic and bioinformatic approaches suggested that the genetic determination of magnetosome formation is complex and may potentially involve 25 to 50 gene functions (15), with unknown numbers of accessory genes and those controlling signal transduction and motility to achieve effective magnetotaxis (8, 9, 12, 26, 27, 29). However, the functional characterization of these candidate genes has been lagging behind. This is due to technical difficulties and the lack of facile tools for genetic manipulation of MTB. Allelic replacement systems have been established for Magnetospirillum magneticum (18) and Magnetospirillum gryphiswaldense (39, 40), but so far, there are only few examples of these for magnetosome genes that were functionally characterized because of the tedious and cumbersome procedures required for mutant generation (11, 19, 28, 31-32). Most genes controlling magnetosome formation in these and other MTB are located within a genomic magnetosome island (MAI) (34), which is genetically instable during stationary growth (47) and more or less conserved in other MTB (12, 13, 35). Most known magnetosome genes are organized within several conserved operons, which are interspersed with large, poorly conserved genome sections of unknown functions that have been speculated to represent genetic junk irrelevant for magnetotaxis but to cause genetic instability by their high content of repeats and transposable elements (34, 47). Thus, for large-scale functional genome analysis and rearrangements of the MAI, there is a great need for additional and more efficient genetic methods.Artificial genome recombination systems have been described for a number of bacteria. Many of them are based on the Cre-loxP system of the P1 phage (42). The Cre-loxP recombination system is a simple two-component system that is recognized as a powerful genetic tool in a multitude of eukaryotic and prokaryotic organisms (4, 6, 48). The Cre protein belongs to the integrase family of site-specific recombinases and catalyzes reciprocal site-specific recombination of DNA at 34-bp loxP sites, resulting in either excision or inversion, depending on the parallel or antiparallel orientation of the loxP sites, respectively (21). It does not require any host cofactors or accessory proteins (7). Cre-lox deletion has several advantages over other methods, such as a high efficiency and the independency of the length of DNA located between the two lox sites. The utility of Cre-lox systems has been demonstrated in a wide variety of Gram-positive and Gram-negative bacteria (17, 22-23). In several studies, it was applied for the generation of large-scale deletions, as in for example, the Gram-positive Corynebacterium glutamicum (43-46) and Bacillus subtilis (49).In M. gryphiswaldense, the functionality of a Cre-loxP antibiotic marker recycling system (25) has been previously demonstrated by deletion of a single gene based on double-crossover insertion of two loxP sites, followed by subsequent Cre-mediated excision (31). In this study, we describe a novel strategy for Cre-loxP-mediated deletion of large genomic fragments which requires only two single crossovers. The system has been validated by the generation of three large deletions, two single and one combination within the MAI, which demonstrated that the total deleted region of approximately 87 kb is not essential for viability and growth in the laboratory.  相似文献   

4.
Magnetosomes are unique bacterial organelles comprising membrane-enveloped magnetic crystals produced by magnetotactic bacteria. Because of several desirable chemical and physical properties, magnetosomes would be ideal scaffolds on which to display highly complicated biological complexes artificially. As a model experiment for the functional expression of a multisubunit complex on magnetosomes, we examined the display of a chimeric bacterial RNase P enzyme composed of the protein subunit (C5) of Escherichia coli RNase P and the endogenous RNA subunit by expressing a translational fusion of C5 with MamC, a known magnetosome protein, in the magnetotactic bacterium Magnetospirillum gryphiswaldense. As intended, the purified C5 fusion magnetosomes, but not wild-type magnetosomes, showed apparent RNase P activity and the association of a typical bacterial RNase P RNA. Our results demonstrate for the first time that magnetosomes can be employed as scaffolds for the display of multisubunit complexes.Magnetosomes are unique organelles comprising membrane-enveloped magnetic crystals of iron minerals (Fe3O4 or Fe3S4) produced by magnetotactic bacteria (1, 11). The bacteria employ magnetosomes to sense the environmental magnetic field, probably in order to recognize their favorite environments. Compared with chemically or physically synthesized magnetic nanoparticles, magnetosomes have a variety of desirable features, including their genetically controlled uniform size and morphology, characteristic crystal habits, and their coverage by a biological membrane that can be addressed by functionalization (1, 4, 11). Based on these features, magnetosomes would be ideal scaffolds on which to display biological molecules artificially.Until now, several heterologous target proteins have been examined for artificial display on magnetosomes (1, 11). For example, reporter proteins such as luciferase and green fluorescent protein were employed to analyze the targeting, expression, and stability of chimeric proteins displayed on magnetosomes (14, 18, 23, 30, 41). For more-practical applications, general antibody-binding proteins (protein A and protein G) were displayed to capture desired antibodies (16, 17, 25, 33, 34, 37, 41). Such antibody-captured magnetosomes are applicable for the magnetic separation of target molecules and cells. Displays of G protein-coupled receptors (the D1 dopamine receptor and the ligand binding domain of the estrogen receptor) were also examined for screening of drugs targeting these receptors (38, 39, 40).There are two major strategies for the construction of functionalized magnetosomes: subsequent chemical modifications of purified magnetosomes (3) and in vivo expression of modified magnetosome proteins (1, 19). The latter approach is confined to biological molecules that can be expressed as a genetic fusion with a magnetosome protein inside a magnetotactic bacterium. By this approach, the target-displaying magnetosomes can be constructed inside cells or under physiological conditions in the presence of a variety of chaperons, are recoverable under mild conditions employing a magnetic field, and provide control by genetic means. Thus, the approach is highly promising for the display of a naïve target such as a multisubunit complex. To date, however, experimental evidence that magnetosomes can be employed as scaffolds for the display of such targets is still lacking. In order to demonstrate this potential of magnetosomes, here, we examined the display of a holoenzyme of bacterial RNase P, one of the simplest complexes composed of a single RNA and a single protein subunit (10, 12), by expressing a fusion of a protein component of the RNase P and a magnetosome membrane protein.  相似文献   

5.
6.
In the present study, we investigated a group of uncultivated magnetotactic cocci, which was magnetically isolated from a freshwater pond in Beijing, China. Light and transmission electron microscopy showed that these cocci ranged from 1.5 to 2.5 μm and contained two to four chains of magnetite magnetosomes, which sometimes were partially disorganized. Overall, the size of the disorganized magnetosomes was significantly smaller than that arranged in chains. All characterized magnetosome crystals were elongated (shape factor = 0.64) and fall into the single-domain size range (30 to 115 nm). Comparative 16S rRNA gene sequence analysis and fluorescence in situ hybridization showed that the enriched bacteria were a virtually homogeneous population and represented a novel lineage in the Alphaproteobacteria. The closest cultivated relative was magnetotactic coccoid strain MC-1 (88% sequence identity). First-order reversal curve diagrams revealed that these cocci had relatively strong magnetic interactions compared to the single-chain magnetotactic bacteria. Low-temperature magnetic measurements showed that the Verwey transition of them was ∼108 K, confirming magnetite magnetosomes, and the delta ratio δFCZFC was >2. Based on the structure, phylogenetic position and magnetic properties, the enriched magnetotactic cocci of Alphaproteobacteria are provisionally named as “Candidatus Magnetococcus yuandaducum.”Magnetotactic bacteria (MTB) can mineralize intracellular nanosized iron oxides or sulfides called magnetosomes, which in most MTB are normally single-domain (SD) magnetite with a narrow range of grain sizes from 30 to 120 nm (3). The chain configuration of magnetosomes renders MTB able to navigate the oxic-anoxic interface in chemically stratified environments by swimming along the Earth''s magnetic field (13). Diverse MTB, including coccoid, spirillar, rod-shaped bacteria and multicellular magnetotactic prokaryotes with one, two, or more chains of magnetosomes, thrive in a broad range of aquatic environments, which sometimes even are dominant strains of the microbial biomass in sediment (10, 47). Based on their phylogeny, all currently known MTB can be divided into two taxonomic groups: Proteobacteria and Nitrospira phyla (2).When MTB die, the magnetosomes can be preserved in sediment as fossil magnetosomes (or magnetofossils) (6). Fossil magnetosomes have been found in lacustrine and deep-sea sediments (6, 35, 43, 51), which are stable carriers of natural remanence and may play substantial contributions to the bulk magnetization of sediments due to their SD sizes (6, 19, 26, 32, 33). Moreover, since most known MTB are microaerophilic or anaerobic and are concentrated in the oxic-anoxic transition zone, the presence and characteristics of MTB species in vertically stratified sediments can be used as a potential paleoenvironmental proxy (19, 44, 45). However, how to identify bacterial magnetite or greigite (Fe3S4) in sediments is still challenging. Recently, characterizing the magnetic properties of MTB has attracted increasing interests because magnetic techniques are fast and effective in distinguishing bacterial crystals from abiogenic magnetic minerals in sediments (11, 26, 27, 33, 38).In spite of their wide distribution and abundance in aquatic environments, most MTB are intractable, and so far only a few of them, e.g., Magnetospirillum gryphiswaldense strain MSR-1 (41), M. magnetotacticum strain AMB-1 (17), M. magnetotacticum strain MS-1 (4), and magnetotactic coccoid strain MC-1 (24), can be grown in pure culture. Until recently, most insights into the molecular characterizations and magnetic properties of MTB have been based on pure cultures, which have a single magnetosome chain per cell (10, 11, 19, 20, 25, 27, 28, 37, 38, 42, 52). However, knowledge of uncultivated MTB, especially strains with multiple chains of magnetosomes that are commonly encountered in natural environments, remains limited. In the present study, we investigated a population of uncultivated magnetotactic cocci with multiple magnetosome chains, which were abundant in the pond in Yuan Dynasty Capital City Wall Relics Park (Yuandadu Park) in Beijing, China, in order to characterize their morphological features, phylogenetic positions, and magnetic properties and finally to classify them in a provisional Candidatus taxon.  相似文献   

7.
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   

8.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  相似文献   

9.
10.
11.
Most research on growing bacterial colonies on agar plates has concerned the effect of genetic or morphotype variation. Some studies have indicated that there is a correlation between microscopic bacterial motion and macroscopic colonial expansion, especially for swarming strains, but no measurements have been obtained for a single strain to relate the microscopic scale to the macroscopic scale. We examined here a single strain (Paenibacillus dendritiformis type T; tip splitting) to determine both the macroscopic growth of colonies and the microscopic bacterial motion within the colonies. Our multiscale measurements for a variety of growth conditions revealed that motion on the microscopic scale and colonial growth are largely independent. Instead, the growth of the colony is strongly affected by the availability of a surfactant that reduces surface tension.Bacteria are able to colonize many different surfaces through collective behavior such as swarming and biofilm formation. Studies of such behavior (10, 18, 26, 31) have revealed cooperative phenomena on both microscopic and colonial scales (4, 5, 7, 8, 20), including production of extracellular “lubricant-wetting” fluid for movement on medium and hard surfaces (19, 22, 25), chemical signaling such as quorum sensing and chemotactic signaling (1, 12, 27), and the secretion of inhibiting and killing factors (2, 9, 11, 14, 15, 17).Research has suggested possible links between the microscopic behavior of a colony and the rate at which the colony expands (12, 23, 24, 29). For Pseudomonas aeruginosa, increased reversal rates for flagella lead to hyperswarming (a larger colony) (26). Similar flagellar modulation affects Escherichia coli (32); if the bacteria never tumble (flagella rotate only counterclockwise) or only tumble (flagella rotate only clockwise), the final colony is much smaller than a colony formed when the bacteria both swim and tumble. For Rhizobium etli, a correlation has been observed between microscopic swarming motion and expansion of the colony, and an acylhomoserine lactone molecule has been found to be a swarming regulator, as well as a biosurfactant that controls surface activity (12). These studies suggest that there is a correlation between microscopic activity and colonial expansion; however, a mutation may be pleiotropic, affecting both motility and surfactant production. Further, there may be additional, unidentified differences between mutant and wild-type strains. For example, the failure of Bacillus subtilis laboratory strains to swarm is caused by a mutation in a gene (sfp) needed for surfactin synthesis and a mutation(s) in an additional unknown gene(s) (21). Experiments that avoid this ambiguity by studying the response of a single strain exposed to changing physical environments have not been performed. Further, except for measurements of the size of an expanding colony as a function of time (3, 6), no detailed time development studies of a growing bacterial colony have been reported.Here we exposed a single bacterial strain, Paenibacillus dendritiformis type T (tip splitting) (4), to different substrate hardnesses, nutrition levels, and surfactant concentrations to identify the parameters that determine colonial growth. P. dendritiformis is a gram-positive rod-shaped (4 μm by 1 μm) bacterium that swims on top of an agar gel in a thin layer (a few micrometers thick) of fluid, presumably secreted by the bacterial cells. The bacteria develop complex colonial (bush-like) branching patterns that are sensitive to small changes in the environment when the bacteria are grown on nutrient-limited surfaces (low peptone levels [approximately 1 g/liter]) (6). The colonies grow slowly (0.1 mm/h) so the microscopic motion can be followed with a microscope for about 10 min without moving the field of view. Also, this strain shows swarming-like microscopic motion where the bacteria move collectively in whirls and jets. This makes this bacterium well suited for studying simultaneously the development of a colony and the internal structure of branches. We constructed a novel setup to observe 10 growing P. dendritiformis colonies in each experiment, and complementary microscopic measurements were obtained for the velocity field of individual bacteria or small groups of cells within the colonies. Specifically, we measured the “bacterial speed,” which was the average of the values for the velocity vectors for the bacteria in a region near the edge of a growing colony, and the “tip velocity,” which was the speed of the moving growth front at the edge of a colony. We also quantified the collective bacterial motion within the colonies by computing spatial and temporal velocity autocorrelation functions.  相似文献   

12.
The sequestration of iron by mammalian hosts represents a significant obstacle to the establishment of a bacterial infection. In response, pathogenic bacteria have evolved mechanisms to acquire iron from host heme. Bacillus anthracis, the causative agent of anthrax, utilizes secreted hemophores to scavenge heme from host hemoglobin, thereby facilitating iron acquisition from extracellular heme pools and delivery to iron-regulated surface determinant (Isd) proteins covalently attached to the cell wall. However, several Gram-positive pathogens, including B. anthracis, contain genes that encode near iron transporter (NEAT) proteins that are genomically distant from the genetically linked Isd locus. NEAT domains are protein modules that partake in several functions related to heme transport, including binding heme and hemoglobin. This finding raises interesting questions concerning the relative role of these NEAT proteins, relative to hemophores and the Isd system, in iron uptake. Here, we present evidence that a B. anthracis S-layer homology (SLH) protein harboring a NEAT domain binds and directionally transfers heme to the Isd system via the cell wall protein IsdC. This finding suggests that the Isd system can receive heme from multiple inputs and may reflect an adaptation of B. anthracis to changing iron reservoirs during an infection. Understanding the mechanism of heme uptake in pathogenic bacteria is important for the development of novel therapeutics to prevent and treat bacterial infections.Pathogenic bacteria need to acquire iron to survive in mammalian hosts (12). However, the host sequesters most iron in the porphyrin heme, and heme itself is often bound to proteins such as hemoglobin (14, 28, 85). Circulating hemoglobin can serve as a source of heme-iron for replicating bacteria in infected hosts, but the precise mechanisms of heme extraction, transport, and assimilation remain unclear (25, 46, 79, 86). An understanding of how bacterial pathogens import heme will lead to the development of new anti-infectives that inhibit heme uptake, thereby preventing or treating infections caused by these bacteria (47, 68).The mechanisms of transport of biological molecules into a bacterial cell are influenced by the compositional, structural, and topological makeup of the cell envelope. Gram-negative bacteria utilize specific proteins to transport heme through the outer membrane, periplasm, and inner membrane (83, 84). Instead of an outer membrane and periplasm, Gram-positive bacteria contain a thick cell wall (59, 60). Proteins covalently anchored to the cell wall provide a functional link between extracellular heme reservoirs and intracellular iron utilization pathways (46). In addition, several Gram-positive and Gram-negative bacterial genera also contain an outermost structure termed the S (surface)-layer (75). The S-layer is a crystalline array of protein that surrounds the bacterial cell and may serve a multitude of functions, including maintenance of cell architecture and protection from host immune components (6, 7, 18, 19, 56). In bacterial pathogens that manifest an S-layer, the “force field” function of this structure raises questions concerning how small molecules such as heme can be successfully passed from the extracellular milieu to cell wall proteins for delivery into the cell cytoplasm.Bacillus anthracis is a Gram-positive, spore-forming bacterium that is the etiological agent of anthrax disease (30, 33). The life cycle of B. anthracis begins after a phagocytosed spore germinates into a vegetative cell inside a mammalian host (2, 40, 69, 78). Virulence determinants produced by the vegetative cells facilitate bacterial growth, dissemination to major organ systems, and eventually host death (76-78). The release of aerosolized spores into areas with large concentrations of people is a serious public health concern (30).Heme acquisition in B. anthracis is mediated by the action of IsdX1 and IsdX2, two extracellular hemophores that extract heme from host hemoglobin and deliver the iron-porphyrin to cell wall-localized IsdC (21, 45). Both IsdX1 and IsdX2 harbor near iron transporter domains (NEATs), a conserved protein module found in Gram-positive bacteria that mediates heme uptake from hemoglobin and contributes to bacterial pathogenesis upon infection (3, 8, 21, 31, 44, 46, 49, 50, 67, 81, 86). Hypothesizing that B. anthracis may contain additional mechanisms for heme transport, we provide evidence that B. anthracis S-layer protein K (BslK), an S-layer homology (SLH) and NEAT protein (32, 43), is surface localized and binds and transfers heme to IsdC in a rapid, contact-dependent manner. These results suggest that the Isd system is not a self-contained conduit for heme trafficking and imply that there is functional cross talk between differentially localized NEAT proteins to promote heme uptake during infection.  相似文献   

13.
The essential cell division protein FtsL is a substrate of the intramembrane protease RasP. Using heterologous coexpression experiments, we show here that the division protein DivIC stabilizes FtsL against RasP cleavage. Degradation seems to be initiated upon accessibility of a cytosolic substrate recognition motif.Cell division in bacteria is a highly regulated process (1). The division site selection as well as assembly and disassembly of the divisome have to be strictly controlled (1, 4). Although the spatial control of the divisome is relatively well understood (2, 4, 14, 17), mechanisms governing the temporal control of division are still mainly elusive. Regulatory proteolysis was thought to be a potential modulatory mechanism (8, 9). The highly unstable division protein FtsL was shown to be rate limiting for division and would make an ideal candidate for a regulatory factor in the timing of bacterial cell division (7, 9). In Bacillus subtilis, FtsL is an essential protein of the membrane part of the divisome (5, 7, 8). It is necessary for the assembly of the membrane-spanning division proteins, and a knockout is lethal (8, 9, 12). We have previously reported that FtsL is a substrate of the intramembrane protease RasP (5).These findings raised the question of whether RasP can regulate cell division by cleaving FtsL from the division complex. In order to mimic the situation in which FtsL is bound to at least one of its interaction partners, we used a heterologous coexpression system in which we synthesized FtsL and DivIC. It has been reported before that DivIC and FtsL are intimate binding partners in various organisms (6, 9, 15, 21, 22, 26) and that FtsL and DivIC (together with DivIB) can form complexes even in the absence of the other divisome components (6, 21). We therefore asked whether RasP is able to cleave FtsL in the presence of its major interaction partner DivIC, which would argue for the possibility that RasP could cleave FtsL within a mature divisome. In contrast, if interaction with DivIC could stabilize FtsL against RasP cleavage, this result would bring such a model into question. An alternative option for the role of RasP might be the removal of FtsL from the membrane. It has been shown that divisome disassembly and prevention of reassembly are crucial to prevent minicell formation close to the new cell poles (3, 16).  相似文献   

14.
15.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

16.
17.
Understanding the mechanisms underlying potential altered susceptibility to human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) individuals and the later clinical consequences of breakthrough infection can provide insight into strategies to control HIV-1 with an effective vaccine. From our Seattle ES cohort, we identified one individual (LSC63) who seroconverted after over 2 years of repeated unprotected sexual contact with his HIV-1-infected partner (P63) and other sexual partners of unknown HIV-1 serostatus. The HIV-1 variants infecting LSC63 were genetically unrelated to those sequenced from P63. This may not be surprising, since viral load measurements in P63 were repeatedly below 50 copies/ml, making him an unlikely transmitter. However, broad HIV-1-specific cytotoxic T-lymphocyte (CTL) responses were detected in LSC63 before seroconversion. Compared to those detected after seroconversion, these responses were of lower magnitude and half of them targeted different regions of the viral proteome. Strong HLA-B27-restricted CTLs, which have been associated with disease control, were detected in LSC63 after but not before seroconversion. Furthermore, for the majority of the protein-coding regions of the HIV-1 variants in LSC63 (except gp41, nef, and the 3′ half of pol), the genetic distances between the infecting viruses and the viruses to which he was exposed through P63 (termed the exposed virus) were comparable to the distances between random subtype B HIV-1 sequences and the exposed viruses. These results suggest that broad preinfection immune responses were not able to prevent the acquisition of HIV-1 infection in LSC63, even though the infecting viruses were not particularly distant from the viruses that may have elicited these responses.Understanding the mechanisms of altered susceptibility or control of human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) persons may provide invaluable information aiding the design of HIV-1 vaccines and therapy (9, 14, 15, 33, 45, 57, 58). In a cohort of female commercial sex workers in Nairobi, Kenya, a small proportion of individuals remained seronegative for over 3 years despite the continued practice of unprotected sex (12, 28, 55, 56). Similarly, resistance to HIV-1 infection has been reported in homosexual men who frequently practiced unprotected sex with infected partners (1, 15, 17, 21, 61). Multiple factors have been associated with the resistance to HIV-1 infection in ES individuals (32), including host genetic factors (8, 16, 20, 37-39, 44, 46, 47, 49, 59, 63), such as certain HLA class I and II alleles (41), as well as cellular (1, 15, 26, 55, 56), humoral (25, 29), and innate immune responses (22, 35).Seroconversion in previously HIV-resistant Nairobi female commercial sex workers, despite preexisting HIV-specific cytotoxic T-lymphocyte (CTL) responses, has been reported (27). Similarly, 13 of 125 ES enrollees in our Seattle ES cohort (1, 15, 17) have become late seroconverters (H. Zhu, T. Andrus, Y. Liu, and T. Zhu, unpublished observations). Here, we analyze the virology, genetics, and immune responses of HIV-1 infection in one of the later seroconverting subjects, LSC63, who had developed broad CTL responses before seroconversion.  相似文献   

18.
19.
Clade B of the New World arenaviruses contains both pathogenic and nonpathogenic members, whose surface glycoproteins (GPs) are characterized by different abilities to use the human transferrin receptor type 1 (hTfR1) protein as a receptor. Using closely related pairs of pathogenic and nonpathogenic viruses, we investigated the determinants of the GP1 subunit that confer these different characteristics. We identified a central region (residues 85 to 221) in the Guanarito virus GP1 that was sufficient to interact with hTfR1, with residues 159 to 221 being essential. The recently solved structure of part of the Machupo virus GP1 suggests an explanation for these requirements.Arenaviruses are bisegmented, single-stranded RNA viruses that use an ambisense coding strategy to express four proteins: NP (nucleoprotein), Z (matrix protein), L (polymerase), and GP (glycoprotein). The viral GP is sufficient to direct entry into host cells, and retroviral vectors pseudotyped with GP recapitulate the entry pathway of these viruses (5, 13, 24, 31). GP is a class I fusion protein comprising two subunits, GP1 and GP2, cleaved from the precursor protein GPC (4, 14, 16, 18, 21). GP1 contains the receptor binding domain (19, 28), while GP2 contains structural elements characteristic of viral membrane fusion proteins (8, 18, 20, 38). The N-terminal stable signal peptide (SSP) remains associated with the mature glycoprotein after cleavage (2, 39) and plays a role in transport, maturation, and pH-dependent fusion (17, 35, 36, 37).The New World arenaviruses are divided into clades A, B, and C based on phylogenetic relatedness (7, 9, 11). Clade B contains the human pathogenic viruses Junin (JUNV), Machupo (MACV), Guanarito (GTOV), Sabia, and Chapare, which cause severe hemorrhagic fevers in South America (1, 10, 15, 26, 34). Clade B also contains the nonpathogenic viruses Amapari (AMAV), Cupixi, and Tacaribe (TCRV), although mild disease has been reported for a laboratory worker infected with TCRV (29).Studies with both viruses and GP-pseudotyped retroviral vectors have shown that the pathogenic clade B arenaviruses use the human transferrin receptor type 1 (hTfR1) to gain entry into human cells (19, 30). In contrast, GPs from nonpathogenic viruses, although capable of using TfR1 orthologs from other species (1), cannot use hTfR1 (1, 19) and instead enter human cells through as-yet-uncharacterized hTfR1-independent pathways (19). In addition, human T-cell lines serve as useful tools to distinguish these GPs, since JUNV, GTOV, and MACV pseudotyped vectors readily transduce CEM cells, while TCRV and AMAV GP vectors do not (27; also unpublished data). These properties of the GPs do not necessarily reflect a tropism of the pathogenic viruses for human T cells, since viral tropism is influenced by many factors and T cells are not a target for JUNV replication in vivo (3, 22, 25).  相似文献   

20.
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   

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