首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
The combinatorial nature of genetic recombination can potentially provide organisms with immediate access to many more positions in sequence space than can be reached by mutation alone. Recombination features particularly prominently in the evolution of a diverse range of viruses. Despite rapid progress having been made in the characterization of discrete recombination events for many species, little is currently known about either gross patterns of recombination across related virus families or the underlying processes that determine genome-wide recombination breakpoint distributions observable in nature. It has been hypothesized that the networks of coevolved molecular interactions that define the epistatic architectures of virus genomes might be damaged by recombination and therefore that selection strongly influences observable recombination patterns. For recombinants to thrive in nature, it is probably important that the portions of their genomes that they have inherited from different parents work well together. Here we describe a comparative analysis of recombination breakpoint distributions within the genomes of diverse single-stranded DNA (ssDNA) virus families. We show that whereas nonrandom breakpoint distributions in ssDNA virus genomes are partially attributable to mechanistic aspects of the recombination process, there is also a significant tendency for recombination breakpoints to fall either outside or on the peripheries of genes. In particular, we found significantly fewer recombination breakpoints within structural protein genes than within other gene types. Collectively, these results imply that natural selection acting against viruses expressing recombinant proteins is a major determinant of nonrandom recombination breakpoint distributions observable in most ssDNA virus families.Genetic recombination is a ubiquitous biological process that is both central to DNA repair pathways (10, 57) and an important evolutionary mechanism. By generating novel combinations of preexisting nucleotide polymorphisms, recombination can potentially accelerate evolution by increasing the population-wide genetic diversity upon which adaptive selection relies. Recombination can paradoxically also prevent the progressive accumulation of harmful mutations within individual genomes (18, 35, 53). Whereas its ability to defend high-fitness genomes from mutational decay possibly underlies the evolutionary value of sexuality in higher organisms, in many microbial species where pseudosexual genetic exchange is permissible among even highly divergent genomes, recombination can enable access to evolutionary innovations that would otherwise be inaccessible by mutation alone.Such interspecies recombination is fairly common in many virus families (8, 17, 27, 44, 82). It is becoming clear, however, that as with mutation events, most recombination events between distantly related genomes are maladaptive (5, 13, 38, 50, 63, 80). As genetic distances between parental genomes increase, so too does the probability of fitness defects in their recombinant offspring (16, 51). The viability of recombinants is apparently largely dependent on how severely recombination disrupts coevolved intragenome interaction networks (16, 32, 51). These networks include interacting nucleotide sequences that form secondary structures, sequence-specific protein-DNA interactions, interprotein interactions, and amino acid-amino acid interactions within protein three-dimensional folds.One virus family where such interaction networks appear to have a large impact on patterns of natural interspecies recombination are the single-stranded DNA (ssDNA) geminiviruses. As with other ssDNA viruses, recombination is very common among the species of this family (62, 84). Partially conserved recombination hot and cold spots have been detected in different genera (39, 81) and are apparently caused by both differential mechanistic predispositions of genome regions to recombination and natural selection disfavoring the survival of recombinants with disrupted intragenome interaction networks (38, 51).Genome organization and rolling circle replication (RCR)—the mechanism by which geminiviruses and many other ssDNA viruses replicate (9, 67, 79; see reference 24 for a review)—seem to have a large influence on basal recombination rates in different parts of geminivirus genomes (20, 33, 39, 61, 81). To initiate RCR, virion-strand ssDNA molecules are converted by host-mediated pathways into double-stranded “replicative-form” (RF) DNAs (34, 67). Initiated by a virus-encoded replication-associated protein (Rep) at a well-defined virion-strand replication origin (v-ori), new virion strands are synthesized on the complementary strand of RF DNAs (28, 73, 74) by host DNA polymerases. Virion-strand replication is concomitant with the displacement of old virion strands, which, once complete, yields covalently closed ssDNA molecules which are either encapsidated or converted into additional RF DNAs. Genome-wide basal recombination rates in ssDNA viruses are probably strongly influenced by the specific characteristics of host DNA polymerases that enable RCR. Interruption of RCR has been implicated directly in geminivirus recombination (40) and is most likely responsible for increased basal recombination rates both within genes transcribed in the opposite direction from that of virion-strand replication (40, 71) and at the v-ori (1, 9, 20, 69, 74).Whereas most ssDNA virus families replicate via either a rolling circle mechanism (the Nanoviridae, Microviridae, and Geminiviridae) (3, 23, 24, 31, 59, 67, 74) or a related rolling hairpin mechanism (the Parvoviridae) (25, 76), among the Circoviridae only the Circovirus genus is known to use RCR (45). Although the Gyrovirus genus (the other member of the Circoviridae) and the anelloviruses (a currently unclassified ssDNA virus group) might also use RCR, it is currently unknown whether they do or not (78). Additionally, some members of the Begomovirus genus of the Geminiviridae either have a second genome component, called DNA-B, or are associated with satellite ssDNA molecules called DNA-1 and DNA-Beta, all of which also replicate by RCR (1, 47, 68).Recombination is known to occur in the parvoviruses (19, 43, 70), microviruses (66), anelloviruses (40, 46), circoviruses (11, 26, 60), nanoviruses (30), geminivirus DNA-B components, and geminivirus satellite molecules (2, 62). Given that most, if not all, of these ssDNA replicons are evolutionarily related to and share many biological features with the geminiviruses (22, 31, 36), it is of interest to determine whether conserved recombination patterns observed in the geminiviruses (61, 81) are evident in these other groups. To date, no comparative analyses have ever been performed with different ssDNA virus families to identify, for example, possible influences of genome organization on recombination breakpoint distributions found in these viruses.Here we compare recombination frequencies and recombination breakpoint distributions in most currently described ssDNA viruses and satellite molecules and identify a number of sequence exchange patterns that are broadly conserved across this entire group.  相似文献   

2.
3.
4.
Synaptonemal complex (SC) proteins Hop1 and Mek1 have been proposed to promote homologous recombination in meiosis of Saccharomyces cerevisiae by establishment of a barrier against sister chromatid recombination. Therefore, it is interesting to know whether the homologous proteins play a similar role in Schizosaccharomyces pombe. Unequal sister chromatid recombination (USCR) was found to be increased in hop1 and mek1 single and double deletion mutants in assays for intrachromosomal recombination (ICR). Meiotic intergenic (crossover) and intragenic (conversion) recombination between homologous chromosomes was reduced. Double-strand break (DSB) levels were also lowered. Notably, deletion of hop1 restored DSB repair in rad50S meiosis. This may indicate altered DSB repair kinetics in hop1 and mek1 deletion strains. A hypothesis is advanced proposing transient inhibition of DSB processing by Hop1 and Mek1 and thus providing more time for repair by interaction with the homologous chromosome. Loss of Hop1 and Mek1 would then result in faster repair and more interaction with the sister chromatid. Thus, in S. pombe meiosis, where an excess of sister Holliday junction over homologous Holliday junction formation has been demonstrated, Hop1 and Mek1 possibly enhance homolog interactions to ensure wild-type level of crossover formation rather than inhibiting sister chromatid interactions.Sexual reproduction in eukaryotes involves formation of haploid gametes from diploid cells by one round of DNA replication, pairing of the homologous chromosomes, and recombination and then by the two meiotic divisions (53). In fungi the gametes differentiate into haploid spores, which germinate to form vegetative cells. Crossover (CO) formation between homologous chromosomes and DNA repair processes between sister chromatids are required for spore viability (10, 55, 58).In vegetative cells homologous recombination (HR) is important for repair of DNA damage and stalled replication forks, with the sister chromatid as the preferred partner (28). Many of the enzymes involved in mitotic HR also contribute to meiotic recombination. In addition, meiosis-specific cytological structures and enzymes enhance recombination frequency (meiotic induction) and shift partner preference from sister chromatids to homologous chromosomes (3, 47, 64, 74). In detail the steps of HR vary between different types of sequence organization (allelic versus sister versus ectopic), between different types of DNA damage, between meiotic and mitotic cells, and between species (10, 55, 58).Meiotic recombination, including CO formation, is initiated by DNA double-strand breaks (DSBs). In Saccharomyces cerevisiae and other eukaryotes, DSBs are formed by Spo11. Many cofactors are required (29). The Schizosaccharomyces pombe homolog is Rec12, also requiring auxiliary factors whose elimination leads to loss of meiotic DSB formation (12). The 5′ single-strand ends at DSBs are processed by nucleases. In S. cerevisiae the MRX complex made up by the proteins Rad50, Mre11, and Xrs2 is required for this resection, as well as for DSB formation. The corresponding MRN complex of S. pombe (Rad50, Rad32, and Nbs1) is not required for DSB formation but is essential for DSB repair (43, 72). Deletion of rad50, rad32, or ctp1 (homologous to SAE2/COM1 in S. cerevisiae and CtIP in humans) leads to very low spore viability. These proteins are also essential for DSB processing (23, 24, 32, 43, 60, 62).Free DNA 3′ ends at DSBs are recruited for invasion of a sister or homologous chromatid by the strand transfer proteins Rad51 and Dmc1, again involving many accessory proteins (16). This results in the central intermediates of HR: heteroduplex DNA consisting of single strands originating from different chromatids and Holliday junctions (HJs). In S. cerevisiae HJs form preferably between homologs with a two- to sixfold excess over intersister HJs (64). Surprisingly, meiotic HJs form with about a fourfold excess between sisters in S. pombe (11). Eventually the intermediates are resolved into crossover (CO) and noncrossover (NCO) events. COs show exchange of the flanking sequences of the two chromatids involved and usually carry a patch of conversion (unilateral transfer of DNA sequences from one chromatid to its interacting partner) near the DSB site. NCOs are conversion events without associated COs (22). In S. pombe loss of core HR functions leads to very low spore viability: deletion of rad51 but not of dmc1 (20), double mutation of rad54 and rdh54 (7), inactivation of the endonuclease activity encoded by mus81 and eme1 (5, 52), and combined deletion of rad22 and rti1 (homologs of RAD52 of S. cerevisiae). But, differently from the other core functions, Rad22 and Rti1 are not required for CO and NCO (50).Early in meiotic prophase of many eukaryotes, axial elements (called lateral elements in later stages) form along sister chromatids, and pairing of homologous chromosomes is initiated, leading to juxtaposition of the homologous chromosomes along their whole length in the synaptonemal complex (SC) (54). In S. pombe no SC is formed, but linear elements (LEs), resembling axial elements of other eukaryotes, are formed. LEs do not form continuously along the chromosomes (1) but load the proteins Rec10, Hop1, and Mek1 (36, 44, 57), which are homologs of, or at least related, to the S. cerevisiae proteins Red1, Hop1, and Mek1, respectively, localizing to axial/lateral elements (2, 67). Hop1 carries a HORMA domain, also present in proteins associating with axial elements and regulating the progress of recombination in higher eukaryotes: Arabidopsis thaliana (61), Caenorhabditis elegans (9, 41), and mammals (18).In S. cerevisiae localization of Hop1 and Mek1 (meiosis-specific protein kinase) to axial elements is dependent on Red1 (2, 67). Mutation of the three S. cerevisiae genes results in reduction of DSB formation, CO and conversion frequencies, and spore viability (26, 31, 59). Direct comparison of unequal sister chromatid recombination (USCR) frequencies in an assay excluding the scoring of intrachromatid recombination (ICR) revealed no increase in the hop1 null mutant but about fourfold increases in the red1 and mek1 null mutants (69). The S. cerevisiae Hop1, Red1, and Mek1 proteins are involved in biasing meiotic DSB repair to occur between homologous chromosomes rather than between sister chromatids (47). Activated Mek1 kinase is required for the inhibition of sister chromatid-mediated DSB repair by Rad51, when the DMC1 gene is deleted and the meiotic recombination checkpoint is activated (4, 27, 38, 47). For Mek1 activation, phosphorylation of Hop1 by the Mec1/Tel1 kinases is also required (6).Less is known about the S. pombe proteins. Hop1 of S. pombe was identified as a nonsignificant hit by sequence comparison with full-length S. cerevisiae Hop1 and contains an N-terminal HORMA domain and a central zinc finger motif like Hop1 in S. cerevisiae. In addition they share a short homology block toward the C terminus (36). The Mek1 protein of S. pombe shares 34% identity and 54% similarity with its S. cerevisiae counterpart along the whole sequence. It contains an FHA domain in the N-terminal part like the other members of its family of checkpoint kinases and is involved in regulation of the meiotic cell cycle (57). Hop1 and Mek1 are strongly expressed in meiosis but not expressed or only slightly expressed in vegetative cells (42, 57). In prophase both proteins localize to LEs as defined by colocalization with the LE component Rec10 (36). Deletion of the distant RED1 homolog rec10 abolishes LE formation (36, 44) and strongly reduces meiotic recombination (17, 70). Rec10, but not Hop1 and Mek1, is required for localization of Rec7 (a distant homolog of S. cerevisiae Rec114) to meiotic chromosomes (34). Rec7 and Rec10 are required for Rec12 activity (12, 29).Obtaining information on the functions of Hop1 and Mek1 in S. pombe was the aim of the work presented here, especially on their possible roles in homolog versus sister discrimination for DSB repair. Deletion mutants have been studied with respect to spore viability and the frequencies of CO and conversion. They have also been assessed for genetic recombination events between sister chromatids in the known PS1 assay (63) and the newly developed VL1 assay (for details, see Fig. Fig.3).3). Physical analysis of DSB formation and repair has been performed in meiotic time course experiments. It is proposed that S. pombe Hop1 and Mek1 are promoting interactions between homologous chromosomes rather than inhibiting interactions between sister chromatids.Open in a separate windowFIG. 3.PS1 and VL1 assay systems for intrachromosomal recombination. Strains with constructs carrying repeated DNA sequences have been assayed for prototroph formation either by intrachromatid recombination (ICR, yielding prototrophs only in PS1) or by unequal sister chromatid recombination (USCR, in PS1 and VL1). Crosses of the constructs were performed with strains carrying a deletion of the ade6 gene to exclude other homologous recombination events. (A) The PS1 assay involves copies of the ade6 gene inactivated by either the hot spot mutation M26 or the mutation 469. The repeated sequences are separated by the ura4+ marker (63). ICR (left) or USCR (right) between the repeated sequences can lead to formation of adenine prototrophs that have lost the ura4+ marker by crossover (CO) or single-strand annealing (SSA) events. Adenine prototrophs maintaining the ura4+ marker can derive from noncrossover (NCO) events. Both types of pairing may lead to CO or NCO products. (B) The newly constructed VL1 assay (see the supplemental material) involves different truncations of the ade6 gene separated by the hygR marker (also called hphMX6), conferring hygromycin resistance. The left truncation carries a 3′ portion of ade6; the right truncation carries a 5′ portion of ade6. While the gray parts of the truncations are not overlapping, the white sections of 500-bp length are of almost identical sequence, allowing for homologous pairing. CO and SSA products resulting from ICR retain only the central portion of ade6 and remain auxotrophic. Adenine prototrophic CO and NCO products resulting from USCR both retain hygromycin resistance. Note that NCO events may arise through loop formation of one sister chromatid and pairing with a single block (500 bp) of the repeated ade6 sequence (39).  相似文献   

5.
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   

6.
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   

7.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

8.
9.
The bacterium Helicobacter pylori is remarkable for its ability to persist in the human stomach for decades without provoking sterilizing immunity. Since repetitive DNA can facilitate adaptive genomic flexibility via increased recombination, insertion, and deletion, we searched the genomes of two H. pylori strains for nucleotide repeats. We discovered a family of genes with extensive repetitive DNA that we have termed the H. pylori RD gene family. Each gene of this family is composed of a conserved 3′ region, a variable mid-region encoding 7 and 11 amino acid repeats, and a 5′ region containing one of two possible alleles. Analysis of five complete genome sequences and PCR genotyping of 42 H. pylori strains revealed extensive variation between strains in the number, location, and arrangement of RD genes. Furthermore, examination of multiple strains isolated from a single subject''s stomach revealed intrahost variation in repeat number and composition. Despite prior evidence that the protein products of this gene family are expressed at the bacterial cell surface, enzyme-linked immunosorbent assay and immunoblot studies revealed no consistent seroreactivity to a recombinant RD protein by H. pylori-positive hosts. The pattern of repeats uncovered in the RD gene family appears to reflect slipped-strand mispairing or domain duplication, allowing for redundancy and subsequent diversity in genotype and phenotype. This novel family of hypervariable genes with conserved, repetitive, and allelic domains may represent an important locus for understanding H. pylori persistence in its natural host.Helicobacter pylori, a gram-negative bacterium, is remarkable for its ability to persist in the human stomach for decades. Colonization with H. pylori increases risk for peptic ulcer disease and gastric adenocarcinoma (53, 70) and elicits a vigorous immune response (15). The persistence of H. pylori occurs in a niche in the human body previously considered inhospitable to microbial colonization: the acidic stomach replete with proteolytic enzymes.H. pylori strains exhibit substantial genetic diversity, including extensive variation in the presence, arrangement, order, and identity of genes (2, 4-7, 25, 51, 74). Furthermore, analyses of multiple single-colony H. pylori isolates from separate stomach biopsy specimens of individual patients have demonstrated diversity, both within hosts (27, 65), and over time (36). The mechanisms that generate H. pylori genetic diversity may be among the factors that enable persistence in this environment (3, 28).While the natural ability of H. pylori for transformation and recombination may explain some of the intra- and interhost genetic variation observed in this bacterium (43), point mutations and interspecies recombination alone are not sufficient for explaining the extent of the variation in H. pylori (14, 32). The initial genomic sequencing of H. pylori strains 26695 and J99 (6, 72) revealed large amounts of repetitive DNA (1, 59). DNA repeats in bacteria are associated with mechanisms of plasticity, such as phase variation (49, 67); slipped-strand mispairing (41, 46); and increased rates of recombination, deletion, and insertion (17, 60, 62). Because many of the recombination repair and mismatch repair mechanisms common in bacteria are absent or modified in H. pylori (28-30, 56, 76), this organism may be particularly susceptible to the diversifying effects of repetitive DNA. In fact, loci in the H. pylori genome containing repetitive DNA have been shown to exhibit extensive inter- and intrahost variation (9, 10, 28, 37).We hypothesized that identification of repetitive DNA hotspots in H. pylori would allow the recognition of genes whose variation could aid in persistence. To examine this hypothesis, we conducted in silico analyses to identify open reading frames (ORFs) enriched for DNA repeats and then used a combination of sequence analyses and immunoassays to examine the patterns associated with the specific repetitive DNA observed. Our approach led to the realization that a previously identified H. pylori-specific gene family (19, 52) exhibits extensive genetic variation at multiple levels.  相似文献   

10.
Sister chromatid recombination (SCR) is a potentially error-free pathway for the repair of DNA lesions associated with replication and is thought to be important for suppressing genomic instability. The mechanisms regulating the initiation and termination of SCR in mammalian cells are poorly understood. Previous work has implicated all the Rad51 paralogs in the initiation of gene conversion and the Rad51C/XRCC3 complex in its termination. Here, we show that hamster cells deficient in the Rad51 paralog XRCC2, a component of the Rad51B/Rad51C/Rad51D/XRCC2 complex, reveal a bias in favor of long-tract gene conversion (LTGC) during SCR. This defect is corrected by expression of wild-type XRCC2 and also by XRCC2 mutants defective in ATP binding and hydrolysis. In contrast, XRCC3-mediated homologous recombination and suppression of LTGC are dependent on ATP binding and hydrolysis. These results reveal an unexpectedly general role for Rad51 paralogs in the control of the termination of gene conversion between sister chromatids.DNA double-strand breaks (DSBs) are potentially dangerous lesions, since their misrepair may cause chromosomal translocations, gene amplifications, loss of heterozygosity (LOH), and other types of genomic instability characteristic of human cancers (7, 9, 21, 40, 76, 79). DSBs are repaired predominantly by nonhomologous end joining or homologous recombination (HR), two evolutionarily conserved DSB repair mechanisms (8, 12, 16, 33, 48, 60, 71). DSBs generated during the S or G2 phase of the cell cycle may be repaired preferentially by HR, using the intact sister chromatid as a template for repair (12, 26, 29, 32, 71). Sister chromatid recombination (SCR) is a potentially error-free pathway for the repair of DSBs, which has led to the proposal that SCR protects against genomic instability, cancer, and aging. Indeed, a number of human cancer predisposition genes are implicated in SCR control (10, 24, 45, 57, 75).HR entails an initial processing of the DSB to generate a free 3′ single-stranded DNA (ssDNA) overhang (25, 48, 56). This is coupled to the loading of Rad51, the eukaryotic homolog of Escherichia coli RecA, which polymerizes to form an ssDNA-Rad51 “presynaptic” nucleoprotein filament. Formation of the presynaptic filament is tightly regulated and requires the concerted action of a large number of gene products (55, 66, 68). Rad51-coated ssDNA engages in a homology search by invading homologous duplex DNA. If sufficient homology exists between the invading and invaded strands, a triple-stranded synapse (D-loop) forms, and the 3′ end of the invading (nascent) strand is extended, using the donor as a template for gene conversion. This recombination intermediate is thought to be channeled into one of the following two major subpathways: classical gap repair or synthesis-dependent strand annealing (SDSA) (48). Gap repair entails the formation of a double Holliday junction, which may resolve into either crossover or noncrossover products. Although this is a major pathway in meiotic recombination, crossing-over is highly suppressed in somatic eukaryotic cells (26, 44, 48). Indeed, the donor DNA molecule is seldom rearranged during somatic HR, suggesting that SDSA is the major pathway for the repair of somatic DSBs (26, 44, 49, 69). SDSA terminates when the nascent strand is displaced from the D-loop and pairs with the second end of the DSB to form a noncrossover product. The mechanisms underlying displacement of the nascent strand are not well understood. However, failure to displace the nascent strand might be expected to result in the production of longer gene conversion tracts during HR (36, 44, 48, 63).Gene conversion triggered in response to a Saccharomyces cerevisiae or mammalian chromosomal DSB generally results in the copying of a short (50- to 300-bp) stretch of information from the donor (short-tract gene conversion [STGC]) (14, 47, 48, 67, 69). A minority of gene conversions in mammalian cells entail more-extensive copying, generating gene conversion tracts that are up to several kilobases in length (long-tract gene conversion [LTGC]) (26, 44, 51, 54, 64). In yeast, very long gene conversions can result from break-induced replication (BIR), a highly processive form of gene conversion in which a bona fide replication fork is thought to be established at the recombination synapse (11, 36, 37, 39, 61, 63). In contrast, SDSA does not require lagging-strand polymerases and appears to be much less processive than a conventional replication fork (37, 42, 78). BIR in yeast has been proposed to play a role in LOH in aging yeast, telomere maintenance, and palindromic gene amplification (5, 41, 52). It is unclear to what extent a BIR-like mechanism operates in mammalian cells, although BIR has been invoked to explain telomere elongation in tumors lacking telomerase (13). It is currently unknown whether LTGC and STGC in somatic mammalian cells are products of mechanistically distinct pathways or whether they represent alternative outcomes of a common SDSA pathway.Vertebrate cells contain five Rad51 paralogs—polypeptides with limited sequence homology to Rad51—Rad51B, Rad51C, Rad51D, XRCC2, and XRCC3 (74). The Rad51 paralogs form the following two major complexes: Rad51B/Rad51C/Rad51D/XRCC2 (BCDX2) and Rad51C/XRCC3 (CX3) (38, 73). Genetic deletion of any one of the rad51 paralogs in the mouse germ line produces early embryonic lethality, and mouse or chicken cells lacking any of the rad51 paralogs reveal hypersensitivity to DNA-damaging agents, reduced frequencies of HR and of sister chromatid exchanges, increased chromatid-type errors, and defective sister chromatid cohesion (18, 72, 73, 82). Collectively, these data implicate the Rad51 paralogs in SCR regulation. The purified Rad51B/Rad51C complex has been shown to assist Rad51-mediated strand exchange (62). XRCC3 null or Rad51C null hamster cells reveal a bias toward production of longer gene conversion tracts, suggesting a role for the CX3 complex in late stages of SDSA (6, 44). Rad51C copurifies with branch migration and Holliday junction resolution activities in mammalian cell extracts (35), and XRCC3, but not XRCC2, facilitates telomere shortening by reciprocal crossing-over in telomeric T loops (77). These data, taken together with the meiotic defects observed in Rad51C hypomorphic mice, suggest a specialized role for CX3, but not for BCDX2, in resolving Holliday junction structures (31, 58).To further address the roles of Rad51 paralogs in late stages of recombination, we have studied the balance between long-tract (>1-kb) and short-tract (<1-kb) SCR in XRCC2 mutant hamster cells. We found that DSB-induced gene conversion in both XRCC2 and XRCC3 mutant cells is biased in favor of LTGC. These defects were suppressed by expression of wild-type (wt) XRCC2 or XRCC3, respectively, although the dependence upon ATP binding and hydrolysis differed between the two Rad51 paralogs. These results indicate that Rad51 paralogs play a more general role in determining the balance between STGC and LTGC than was previously appreciated and suggest roles for both the BCDX2 and CX3 complexes in influencing the termination of gene conversion in mammals.  相似文献   

11.
12.
Rousettus bat coronavirus HKU9 (Ro-BatCoV HKU9), a recently identified coronavirus of novel Betacoronavirus subgroup D, from Leschenault''s rousette, was previously found to display marked sequence polymorphism among genomes of four strains. Among 10 bats with complete RNA-dependent RNA polymerase (RdRp), spike (S), and nucleocapsid (N) genes sequenced, three and two sequence clades for all three genes were codetected in two and five bats, respectively, suggesting the coexistence of two or three distinct genotypes of Ro-BatCoV HKU9 in the same bat. Complete genome sequencing of the distinct genotypes from two bats, using degenerate/genome-specific primers with overlapping sequences confirmed by specific PCR, supported the coexistence of at least two distinct genomes in each bat. Recombination analysis using eight Ro-BatCoV HKU9 genomes showed possible recombination events between strains from different bat individuals, which may have allowed for the generation of different genotypes. Western blot assays using recombinant N proteins of Ro-BatCoV HKU9, Betacoronavirus subgroup A (HCoV-HKU1), subgroup B (SARSr-Rh-BatCoV), and subgroup C (Ty-BatCoV HKU4 and Pi-BatCoV HKU5) coronaviruses were subgroup specific, supporting their classification as separate subgroups under Betacoronavirus. Antibodies were detected in 75 (43%) of 175 and 224 (64%) of 350 tested serum samples from Leschenault''s rousette bats by Ro-BatCoV HKU9 N-protein-based Western blot and enzyme immunoassays, respectively. This is the first report describing coinfection of different coronavirus genotypes in bats and coronavirus genotypes of diverse nucleotide variation in the same host. Such unique phenomena, and the unusual instability of ORF7a, are likely due to recombination which may have been facilitated by the dense roosting behavior and long foraging range of Leschenault''s rousette.Coronaviruses infect a wide variety of animals in which they can cause respiratory, enteric, hepatic, and neurological diseases of various severities. Based on genotypic and serological characterization, coronaviruses were traditionally classified into three distinct groups, groups 1, 2, and 3 (3, 27, 59). Recently, the Coronavirus Study Group of the International Committee for Taxonomy of Viruses has renamed the traditional group 1, 2, and 3 coronaviruses as Alphacoronavirus, Betacoronavirus, and Gammacoronavirus, respectively (http://talk.ictvonline.org/media/p/1230.aspx). Coronaviruses are known to have a high frequency of recombination as a result of their unique mechanism of viral replication (27). Such tendency for recombination and high mutation rates may allow them to adapt to new hosts and ecological niches (24, 47, 52).The severe acute respiratory syndrome (SARS) epidemic has boosted interest in the study of coronaviruses in humans and animals (21, 34, 38, 41, 54). In the past few years, there has been a dramatic increase in the number of newly described human and animal coronaviruses (2, 4, 5, 8-10, 15-20, 23, 25, 28, 30, 32, 35, 36, 39, 43, 45, 50, 51, 53, 56, 58). Two novel human coronaviruses, human coronavirus NL63 (HCoV-NL63) and human coronavirus HKU1 (HCoV-HKU1), belonging to Alphacoronavirus and Betacoronavirus, respectively, have been discovered, in addition to the human coronavirus OC43 (HCoV-OC43), human coronavirus 229E (HCoV-229E), and SARS coronavirus (SARS-CoV) (17, 29, 45, 53, 55). We have also previously described the discovery of a diversity of novel coronaviruses in wild bats and birds in China, including SARSr-Rh-BatCoV, belonging to Betacoronavirus subgroup B, from Chinese horseshoe bats (30, 48, 56). Among these novel coronaviruses, three avian coronaviruses were found to belong to a novel subgroup of Gammacoronavirus (Gammacoronavirus subgroup C), while three bat coronaviruses were found to belong to two novel subgroups of Betacoronavirus (Betacoronavirus subgroups C and D) (48, 50). Based on the presence of the huge diversity of coronaviruses in Alphacoronavirus and Betacoronavirus among various bat species, bats are likely the reservoir for the ancestor of these two coronavirus genera (47).During our genome analysis of these novel coronaviruses, one of them, Rousettus bat coronavirus HKU9 (Ro-BatCoV HKU9), belonging to Betacoronavirus subgroup D, which was identified in Leschenault''s rousette bats, was found to display marked nucleotide and amino acid sequence polymorphism among the four strains with complete genome sequences (50). In our study on HCoV-HKU1, it has been shown that such sequence polymorphisms may indicate the presence of different genotypes (52). By complete genome sequence analysis of the potentially different genotypes of HCoV-HKU1, we have demonstrated for the first time natural recombination in a human coronavirus, resulting in the generation of at least three genotypes (52). We have also recently shown that recombination between different strains of SARSr-Rh-BatCoV from different regions of China may have given rise to the emergence of civet SARSr-CoV (31). To investigate the presence of different genotypes of Ro-BatCoV HKU9, the complete RNA-dependent RNA polymerase (RdRp) (corresponding to nsp12), spike (S), and nucleocapsid (N) gene sequences of Ro-BatCoV HKU9 from 10 additional bats were determined. Since sequence analysis showed the possible coexistence of different genotypes in seven bat individuals, complete genome sequencing of these distinct genotypes from two bats was carried out to investigate for possible recombination events among the different genotypes. In addition, serological characterization of Ro-BatCoV HKU9 was also performed by Western blot and enzyme immunoassays using recombinant Ro-BatCoV HKU9 nucleocapsid proteins and recombinant nucleocapsid proteins of Betacoronavirus subgroup A, B, and C coronaviruses to determine possible cross-reactivity among the different Betacoronavirus subgroups and the seroepidemiology of Ro-BatCoV HKU9 in Leschenault''s rousette bats.  相似文献   

13.
14.
Using both sequence- and function-based metagenomic approaches, multiple antibiotic resistance determinants were identified within metagenomic libraries constructed from DNA extracted from bacterial chromosomes, plasmids, or viruses within an activated sludge microbial assemblage. Metagenomic clones and a plasmid that in Escherichia coli expressed resistance to chloramphenicol, ampicillin, or kanamycin were isolated, with many cloned DNA sequences lacking any significant homology to known antibiotic resistance determinants.Activated sludge in wastewater treatment plants is an open system with a dynamic and phylogenetically diverse microbial community (2, 3, 6, 7, 10, 11). Since the activated sludge process promotes cellular interactions among diverse microorganisms, there is great potential for the lateral transfer of antibiotic resistance genes between microbes in activated sludge and in downstream environments. Several studies have previously identified antibiotic resistance determinants from wastewater communities that are carried on bacterial chromosomes (1, 4, 14) and plasmids (9, 12, 13), but to our knowledge, a simultaneous metagenomic survey of antibiotic resistance determinants from all three genetic reservoirs (i.e., chromosomes, plasmids, and viruses) has never been performed within the same environment. To achieve a more comprehensive assessment of antibiotic resistance genes in the activated sludge microbial community, this study used both function- and sequence-based metagenomic approaches to identify antibiotic resistance determinants carried on bacterial chromosomes, plasmids, or viruses within an activated sludge microbial assemblage.  相似文献   

15.
The tomato pathotype of Alternaria alternata produces host-specific AAL toxin and causes Alternaria stem canker on tomato. A polyketide synthetase (PKS) gene, ALT1, which is involved in AAL toxin biosynthesis, resides on a 1.0-Mb conditionally dispensable chromosome (CDC) found only in the pathogenic and AAL toxin-producing strains. Genomic sequences of ALT1 and another PKS gene, both of which reside on the CDC in the tomato pathotype strains, were compared to those of tomato pathotype strains collected worldwide. This revealed that the sequences of both CDC genes were identical among five A. alternata tomato pathotype strains having different geographical origins. On the other hand, the sequences of other genes located on chromosomes other than the CDC are not identical in each strain, indicating that the origin of the CDC might be different from that of other chromosomes in the tomato pathotype. Telomere fingerprinting and restriction fragment length polymorphism analyses of the A. alternata strains also indicated that the CDCs in the tomato pathotype strains were identical, although the genetic backgrounds of the strains differed. A hybrid strain between two different pathotypes was shown to harbor the CDCs derived from both parental strains with an expanded range of pathogenicity, indicating that CDCs can be transmitted from one strain to another and stably maintained in the new genome. We propose a hypothesis whereby the ability to produce AAL toxin and to infect a plant could potentially be distributed among A. alternata strains by horizontal transfer of an entire pathogenicity chromosome. This could provide a possible mechanism by which new pathogens arise in nature.Fungi produce a huge variety of secondary metabolites. Some plant-pathogenic fungi, especially necrotrophic pathogens that kill plant cells during invasion, produce phytotoxic metabolites to impair host tissue functions (20, 30, 42, 47). Phytotoxins produced by fungal plant pathogens are generally low-molecular-weight secondary metabolites that exert toxic effects on host plants. Among these phytotoxins, host-specific toxins (HSTs) are critical determinants of pathogenicity or virulence in several plant-pathogen interactions (13, 30, 33, 40, 42, 47, 49).Recent advances in molecular biological techniques for fungi have led to the identification of fungal genes involved in pathogenesis, as exemplified by those used in the biosynthesis of toxic secondary metabolites, such as HSTs. Genes involved in the biosynthesis of secondary metabolites are typically clustered in filamentous fungi, including plant pathogens (20, 24, 44). The origins and evolutionary processes of these gene clusters, however, are largely unknown. Analysis of the arrangement and sequences of genes in the clusters would shed light on how the clusters themselves and their ability to produce toxic secondary metabolites evolved (20, 24, 44).The involvement of horizontal gene transfer (HGT) in the evolution of fungal secondary-metabolite gene clusters has been discussed (34, 44). HGT events are well known in prokaryotes (21, 29), and the genomic regions that have undergone HGT are referred to as pathogenicity or genomic islands (7). In prokaryotes, the mechanisms of HGT are also associated with conjugation, transformation, and transduction (21, 29). Although these transfer mechanisms are generally unknown in eukaryotes such as fungi, interspecific transfer of a virulence gene encoding the production of a critical toxin has been reported in Pyrenophora tritici-repentis (14). There is also clear evidence of recent lateral gene transfer of the ToxA gene from Stagonospora nodorum to P. tritici-repentis (14, 30).In Alternaria alternata plant pathogens (37), we have shown that all strains of the A. alternata pathotypes harbor small extra chromosomes of less than 1.7 Mb, whereas nonpathogenic isolates do not have these small chromosomes (5). A cyclic peptide synthetase gene, AMT, which is involved in host-specific AM toxin biosynthesis of the apple pathotype of A. alternata, was located on a small chromosome of 1.1 to 1.7 Mb, depending on the strain (22, 23). The AF toxin biosynthesis gene cluster was also present on a single small chromosome of 1.05 Mb in the strawberry pathotype of A. alternata (18). Based on biological and pathological observations, those small chromosomes were regarded as supernumerary chromosomes, or conditionally dispensable chromosomes (CDCs) (10, 18, 22). Fungal supernumerary chromosomes, which are not important for normal growth but confer advantages for colonizing an ecological niche, such as infecting host plants, are regarded as CDCs (21). The functions and pathological roles of CDCs have been studied in the pea pathogen Nectria haematococca (11, 17, 25, 32, 43, 46).The origin and evolution of CDCs have been intriguing issues in the study of plant-microbe interactions. The supernumerary chromosomes of certain strains of N. haematococca have been suggested to have a different evolutionary history than essential chromosomes (ECs) in the same genome, and they might have been introduced into the genome by horizontal transfer from another strain (10, 12, 36). In Colletotrichum gloeosporioides, the 2-Mb supernumerary chromosome was transferred from a biotype A strain to a vegetative incompatible biotype B strain (19, 31). Transfer of the chromosome, however, did not affect the pathogenicity of the recipient fungus, perhaps because it did not harbor pathogenicity genes (19, 31). These results suggest that supernumerary chromosomes of fungi might have the capacity for horizontal transfer across an incompatibility barrier between two distinct strains.AAL toxins are HSTs produced by the tomato pathotype of A. alternata (synonym A. alternata f. sp. lycopersici, synonym Alternaria arborescens), the causal agent of Alternaria stem canker disease in tomatoes, which causes severe necrosis of susceptible tomato cultivars (15, 26, 35). AAL toxins and fumonisins of the maize pathogen Gibberella moniliformis are structurally related to sphinganine and termed sphinganine-analogue mycotoxins. AAL toxins and fumonisins are sphinganine-analogue mycotoxins, which are toxic to some plant species and mammalian cells (16, 48). They cause apoptosis in susceptible tomato cells and mammalian cells by inhibiting ceramide biosynthesis (9, 41, 45). In the tomato pathotype of A. alternata-tomato interactions, a major factor in pathogenicity is the production of host-specific AAL toxins capable of inducing cell death only in susceptible cultivars (3, 9, 48).In this study, we describe evidence showing that the ability to produce the host-specific AAL toxin and to infect host tomato plants could potentially be distributed among a population of strains of the A. alternata tomato pathotype by horizontal transfer of an entire pathogenicity chromosome of the pathogen.  相似文献   

16.
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   

17.
18.
19.
20.
Adhesive pili on the surface of the serotype M1 Streptococcus pyogenes strain SF370 are composed of a major backbone subunit (Spy0128) and two minor subunits (Spy0125 and Spy0130), joined covalently by a pilin polymerase (Spy0129). Previous studies using recombinant proteins showed that both minor subunits bind to human pharyngeal (Detroit) cells (A. G. Manetti et al., Mol. Microbiol. 64:968-983, 2007), suggesting both may act as pilus-presented adhesins. While confirming these binding properties, studies described here indicate that Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role as a wall linker. Pili were localized predominantly to cell wall fractions of the wild-type S. pyogenes parent strain and a spy0125 deletion mutant. In contrast, they were found almost exclusively in culture supernatants in both spy0130 and srtA deletion mutants, indicating that the housekeeping sortase (SrtA) attaches pili to the cell wall by using Spy0130 as a linker protein. Adhesion assays with antisera specific for individual subunits showed that only anti-rSpy0125 serum inhibited adhesion of wild-type S. pyogenes to human keratinocytes and tonsil epithelium to a significant extent. Spy0125 was localized to the tip of pili, based on a combination of mutant analysis and liquid chromatography-tandem mass spectrometry analysis of purified pili. Assays comparing parent and mutant strains confirmed its role as the adhesin. Unexpectedly, apparent spontaneous cleavage of a labile, proline-rich (8 of 14 residues) sequence separating the N-terminal ∼1/3 and C-terminal ∼2/3 of Spy0125 leads to loss of the N-terminal region, but analysis of internal spy0125 deletion mutants confirmed that this has no significant effect on adhesion.The group A Streptococcus (S. pyogenes) is an exclusively human pathogen that commonly colonizes either the pharynx or skin, where local spread can give rise to various inflammatory conditions such as pharyngitis, tonsillitis, sinusitis, or erysipelas. Although often mild and self-limiting, GAS infections are occasionally very severe and sometimes lead to life-threatening diseases, such as necrotizing fasciitis or streptococcal toxic shock syndrome. A wide variety of cell surface components and extracellular products have been shown or suggested to play important roles in S. pyogenes virulence, including cell surface pili (1, 6, 32). Pili expressed by the serotype M1 S. pyogenes strain SF370 mediate specific adhesion to intact human tonsil epithelia and to primary human keratinocytes, as well as cultured keratinocyte-derived HaCaT cells, but not to Hep-2 or A549 cells (1). They also contribute to adhesion to a human pharyngeal cell line (Detroit cells) and to biofilm formation (29).Over the past 5 years, pili have been discovered on an increasing number of important Gram-positive bacterial pathogens, including Bacillus cereus (4), Bacillus anthracis (4, 5), Corynebacterium diphtheriae (13, 14, 19, 26, 27, 44, 46, 47), Streptococcus agalactiae (7, 23, 38), and Streptococcus pneumoniae (2, 3, 24, 25, 34), as well as S. pyogenes (1, 29, 32). All these species produce pili that are composed of a single major subunit plus either one or two minor subunits. During assembly, the individual subunits are covalently linked to each other via intermolecular isopeptide bonds, catalyzed by specialized membrane-associated transpeptidases that may be described as pilin polymerases (4, 7, 25, 41, 44, 46). These are related to the classical housekeeping sortase (usually, but not always, designated SrtA) that is responsible for anchoring many proteins to Gram-positive bacterial cell walls (30, 31, 33). The C-terminal ends of sortase target proteins include a cell wall sorting (CWS) motif consisting, in most cases, of Leu-Pro-X-Thr-Gly (LPXTG, where X can be any amino acid) (11, 40). Sortases cleave this substrate between the Thr and Gly residues and produce an intermolecular isopeptide bond linking the Thr to a free amino group provided by a specific target. In attaching proteins to the cell wall, the target amino group is provided by the lipid II peptidoglycan precursor (30, 36, 40). In joining pilus subunits, the target is the ɛ-amino group in the side chain of a specific Lys residue in the second subunit (14, 18, 19). Current models of pilus biogenesis envisage repeated transpeptidation reactions adding additional subunits to the base of the growing pilus, until the terminal subunit is eventually linked covalently via an intermolecular isopeptide bond to the cell wall (28, 41, 45).The major subunit (sometimes called the backbone or shaft subunit) extends along the length of the pilus and appears to play a structural role, while minor subunits have been detected either at the tip, the base, and/or at occasional intervals along the shaft, depending on the species (4, 23, 24, 32, 47). In S. pneumoniae and S. agalactiae one of the minor subunits acts as an adhesin, while the second appears to act as a linker between the base of the assembled pilus and the cell wall (7, 15, 22, 34, 35). It was originally suggested that both minor subunits of C. diphtheriae pili could act as adhesins (27). However, recent data showed one of these has a wall linker role (26, 44) and may therefore not function as an adhesin.S. pyogenes strain SF370 pili are composed of a major (backbone) subunit, termed Spy0128, plus two minor subunits, called Spy0125 and Spy0130 (1, 32). All three are required for efficient adhesion to target cells (1). Studies employing purified recombinant proteins have shown that both of the minor subunits, but not the major subunit, bind to Detroit cells (29), suggesting both might act as pilus-presented adhesins. Here we report studies employing a combination of recombinant proteins, specific antisera, and allelic replacement mutants which show that only Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role in linking pili to the cell wall.  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