首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Recently, a new type of hybrid resulting from the hybridization between Saccharomyces cerevisiae and Saccharomyces kudriavzevii was described. These strains exhibit physiological properties of potential biotechnological interest. A preliminary characterization of these hybrids showed a trend to reduce the S. kudriavzevii fraction of the hybrid genome. We characterized the genomic constitution of several wine S. cerevisiae × S. kudriavzevii strains by using a combined approach based on the restriction fragment length polymorphism analysis of gene regions, comparative genome hybridizations with S. cerevisiae DNA arrays, ploidy analysis, and gene dose determination by quantitative real-time PCR. The high similarity in the genome structures of the S. cerevisiae × S. kudriavzevii hybrids under study indicates that they originated from a single hybridization event. After hybridization, the hybrid genome underwent extensive chromosomal rearrangements, including chromosome losses and the generation of chimeric chromosomes by the nonreciprocal recombination between homeologous chromosomes. These nonreciprocal recombinations between homeologous chromosomes occurred in highly conserved regions, such as Ty long terminal repeats (LTRs), rRNA regions, and conserved protein-coding genes. This study supports the hypothesis that chimeric chromosomes may have been generated by a mechanism similar to the recombination-mediated chromosome loss acting during meiosis in Saccharomyces hybrids. As a result of the selective processes acting during fermentation, hybrid genomes maintained the S. cerevisiae genome but reduced the S. kudriavzevii fraction.The genus Saccharomyces consists of seven biological species: S. arboricolus, S. bayanus, S. cariocanus, S. cerevisiae, S. kudriavzevii, S. mikatae, and S. paradoxus (29, 59) and the partially allotetraploid species S. pastorianus (46, 58).The hybrid species S. pastorianus, restricted to lager brewing environments, arose from two or more natural hybridization events between S. cerevisiae and a S. bayanus-like yeast (7, 16, 28, 46). Recent studies of S. bayanus have also revealed the hybrid nature of certain strains of this species, which has subsequently been subdivided into two groups, S. bayanus var. bayanus, containing a variety of hybrid strains, and S. bayanus var. uvarum, also referred to as S. uvarum, that contains nonhybrid strains (45, 46).New hybrids of other species from the genus Saccharomyces have recently been described. Hybrid yeasts of S. cerevisiae and S. kudriavzevii have been characterized among wine (6, 20, 33) and brewing yeasts (21); even triple hybrids of S. cerevisiae, S. bayanus, and S. kudriavzevii have been identified (20, 41).The first natural Saccharomyces interspecific hybrid identified, the lager brewing yeast S. pastorianus (S. carlsbergensis) (42, 57), has become one of the most investigated types of yeast hybrids. The genome structure of these hybrids has been examined by competitive array comparative genome hybridization (aCGH) (5, 16, 28), complete genome sequencing (28), and PCR-restriction fragment length polymorphism (RFLP) analysis of 48 genes and partial sequences of 16 genes (46). The aCGH analyses of several S. pastorianus strains with S. cerevisiae-only DNA arrays (5, 28) revealed the presence of aneuploidies due to deletions of entire regions of the S. cerevisiae fraction of the hybrid genomes. A recent aCGH analysis of S. pastorianus strains with S. cerevisiae and S. bayanus DNA arrays (16) showed two groups of strains according to their genome structure and composition. These groups arose from two independent hybridization events, and each one is characterized by a reduction and an amplification of the S. cerevisiae genome fraction, respectively.The genetic characterization of the wine S. cerevisiae and S. kudriavzevii hybrids by restriction analysis of five nuclear genes located in different chromosomes, 5.8S-ITS rDNA region and the mitochondrial COX2 gene, revealed the presence of three types of hybrids in Swiss wines, thus indicating the presence of different hybrid genomes (20). In a recent study (21), we identified six new types of S. cerevisiae and S. kudriavzevii hybrids among brewing strains, which were compared to wine hybrids by a genetic characterization based on RFLP analysis of 35 protein-encoding genes. This analysis confirmed the presence of three different genome types among wine hybrids that contain putative chimeric chromosomes, probably generated by a recombination between homeologous chromosomes of different parental origins.The aim of the present study is to investigate the genome composition and structure of wine hybrids of S. cerevisiae and S. kudriavzevii. This has been achieved by a combined approach based on the RFLP analysis of 35 gene regions from our previous study, comparative genome hybridizations using S. cerevisiae DNA macroarrays, a ploidy analysis by flow cytometry, and gene dose determinations by quantitative real-time PCR. This multiple approach allowed us to confirm the presence of chimeric chromosomes and define the mechanisms involved in their origins.  相似文献   

2.
The ability to undergo dramatic morphological changes in response to extrinsic cues is conserved in fungi. We have used the model yeast Schizosaccharomyces pombe to determine which intracellular signal regulates the dimorphic switch from the single-cell yeast form to the filamentous invasive growth form. The S. pombe Asp1 protein, a member of the conserved Vip1 1/3 inositol polyphosphate kinase family, is a key regulator of the morphological switch via the cAMP protein kinase A (PKA) pathway. Lack of a functional Asp1 kinase domain abolishes invasive growth which is monopolar, while an increase in Asp1-generated inositol pyrophosphates (PP) increases the cellular response. Remarkably, the Asp1 kinase activity encoded by the N-terminal part of the protein is regulated negatively by the C-terminal domain of Asp1, which has homology to acid histidine phosphatases. Thus, the fine tuning of the cellular response to environmental cues is modulated by the same protein. As the Saccharomyces cerevisiae Asp1 ortholog is also required for the dimorphic switch in this yeast, we propose that Vip1 family members have a general role in regulating fungal dimorphism.Eucaryotic cells are able to define and maintain a particular cellular organization and thus cellular morphology by executing programs modulated by internal and external signals. For example, signals generated within a cell are required for the selection of the growth zone after cytokinesis in the fission yeast Schizosaccharomyces pombe or the emergence of the bud in Saccharomyces cerevisiae (37, 44, 81). Cellular morphogenesis is also subject to regulation by a wide variety of external signals, such as growth factors, temperature, hormones, nutrient limitation, and cell-cell or cell-substrate contact (13, 34, 66, 75, 81). Both types of signals will lead to the selection of growth zones accompanied by the reorganization of the cytoskeleton.The ability to alter the growth form in response to environmental conditions is an important virulence-associated trait of pathogenic fungi which helps the pathogen to spread in and survive the host''s defense system (7, 32). Alteration of the growth form in response to extrinsic signals is not limited to pathogenic fungi but is also found in the model yeasts S. cerevisiae and S. pombe, in which it appears to represent a foraging response (1, 24).The regulation of polarized growth and the definition of growth zones have been studied extensively with the fission yeast S. pombe. In this cylindrically shaped organism, cell wall biosynthesis is restricted to one or both cell ends in a cell cycle-regulated manner and to the septum during cytokinesis (38). This mode of growth requires the actin cytoskeleton to direct growth and the microtubule cytoskeleton to define the growth sites (60). In interphase cells, microtubules are organized in antiparallel bundles that are aligned along the long axis of the cell and grow from their plus ends toward the cell tips. Upon contact with the cell end, microtubule growth will first pause and then undergo a catastrophic event and microtubule shrinkage (21). This dynamic behavior of the microtubule plus end is regulated by a disparate, conserved, microtubule plus end group of proteins, called the +TIPs. The +TIP complex containing the EB1 family member Mal3 is required for the delivery of the Tea1-Tea4 complex to the cell tip (6, 11, 27, 45, 77). The latter complex docks at the cell end and recruits proteins required for actin nucleation (46, 76). Thus, the intricate cross talk between the actin and the microtubule cytoskeleton at specific intracellular locations is necessary for cell cycle-dependent polarized growth of the fission yeast cell.The intense analysis of polarized growth control in single-celled S. pombe makes this yeast an attractive organism for the identification of key regulatory components of the dimorphic switch. S. pombe multicellular invasive growth has been observed for specific strains under specific conditions, such as nitrogen and ammonium limitation and the presence of excess iron (1, 19, 50, 61).Here, we have identified an evolutionarily conserved key regulator of the S. pombe dimorphic switch, the Asp1 protein. Asp1 belongs to the highly conserved family of Vip1 1/3 inositol polyphosphate kinases, which is one of two families that can generate inositol pyrophosphates (PP) (17, 23, 42, 54). The inositol polyphosphate kinase IP6K family, of which the S. cerevisiae Kcs1 protein is a member, is the “classical” family that can phosphorylate inositol hexakisphosphate (IP6) (70, 71). These enzymes generate a specific PP-IP5 (IP7), which has the pyrophosphate at position 5 of the inositol ring (20, 54). The Vip1 family kinase activity was unmasked in an S. cerevisiae strain with KCS1 and DDP1 deleted (54, 83). The latter gene encodes a nudix hydrolase (14, 68). The mammalian and S. cerevisiae Vip1 proteins phosphorylate the 1/3 position of the inositol ring, generating 1/3 diphosphoinositol pentakisphosphate (42). Both enzyme families collaborate to generate IP8 (17, 23, 42, 54, 57).Two modes of action have been described for the high-energy moiety containing inositol pyrophosphates. First, these molecules can phosphorylate proteins by a nonenzymatic transfer of a phosphate group to specific prephosphorylated serine residues (2, 8, 69). Second, inositol pyrophosphates can regulate protein function by reversible binding to the S. cerevisiae Pho80-Pho85-Pho81 complex (39, 40). This cyclin-cyclin-dependent kinase complex is inactivated by inositol pyrophosphates generated by Vip1 when cells are starved of inorganic phosphate (39, 41, 42).Regulation of phosphate metabolism in S. cerevisiae is one of the few roles specifically attributed to a Vip1 kinase. Further information about the cellular function of this family came from the identification of the S. pombe Vip1 family member Asp1 as a regulator of the actin nucleator Arp2/3 complex (22). The 106-kDa Asp1 cytoplasmic protein, which probably exists as a dimer in vivo, acts as a multicopy suppressor of arp3-c1 mutants (22). Loss of Asp1 results in abnormal cell morphology, defects in polarized growth, and aberrant cortical actin cytoskeleton organization (22).The Vip1 family proteins have a dual domain structure which consists of an N-terminal “rimK”/ATP-grasp superfamily domain found in certain inositol signaling kinases and a C-terminal part with homology to histidine acid phosphatases present in phytase enzymes (28, 53, 54). The N-terminal domain is required and sufficient for Vip1 family kinase activity, and an Asp1 variant with a mutation in a catalytic residue of the kinase domain is unable to suppress mutants of the Arp2/3 complex (17, 23, 54). To date, no function has been described for the C-terminal phosphatase domain, and this domain appears to be catalytically inactive (17, 23, 54).Here we describe a new and conserved role for Vip1 kinases in regulating the dimorphic switch in yeasts. Asp1 kinase activity is essential for cell-cell and cell-substrate adhesion and the ability of S. pombe cells to grow invasively. Interestingly, Asp1 kinase activity is counteracted by the putative phosphatase domain of this protein, a finding that allows us to describe for the first time a function for the C-terminal part of Vip1 proteins.  相似文献   

3.
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   

4.
5.
Cytosolic chaperones are a diverse group of ubiquitous proteins that play central roles in multiple processes within the cell, including protein translation, folding, intracellular trafficking, and quality control. These cellular proteins have also been implicated in the replication of numerous viruses, although the full extent of their involvement in viral replication is unknown. We have previously shown that the heat shock protein 40 (hsp40) chaperone encoded by the yeast YDJ1 gene facilitates RNA replication of flock house virus (FHV), a well-studied and versatile positive-sense RNA model virus. To further explore the roles of chaperones in FHV replication, we examined a panel of 30 yeast strains with single deletions of cytosolic proteins that have known or hypothesized chaperone activity. We found that the majority of cytosolic chaperone deletions had no impact on FHV RNA accumulation, with the notable exception of J-domain-containing hsp40 chaperones, where deletion of APJ1 reduced FHV RNA accumulation by 60%, while deletion of ZUO1, JJJ1, or JJJ2 markedly increased FHV RNA accumulation, by 4- to 40-fold. Further studies using cross complementation and double-deletion strains revealed that the contrasting effects of J domain proteins were reproduced by altering expression of the major cytosolic hsp70s encoded by the SSA and SSB families and were mediated in part by divergent effects on FHV RNA polymerase synthesis. These results identify hsp70 chaperones as critical regulators of FHV RNA replication and indicate that cellular chaperones can have both positive and negative regulatory effects on virus replication.The compact genomes of viruses relative to those of other infectious agents restrict their ability to encode all proteins required to complete their replication cycles. To circumvent this limitation, viruses often utilize cellular factors or processes to complete essential steps in replication. One group of cellular proteins frequently targeted by viruses are cellular chaperones, which include a diverse set of heat shock proteins (hsps) that normally facilitate cellular protein translation, folding, trafficking, and degradation (18, 64). The connection between viruses and cellular chaperones was originally identified in bacteria, where the Escherichia coli hsp40 and hsp70 homologues, encoded by dnaJ and dnaK, respectively, were identified as bacterial genes essential for bacteriophage λ DNA replication (62). Research over the past 30 years has further revealed the importance of cellular chaperones in viral replication, such that the list of virus-hsp connections is now quite extensive and includes viruses from numerous families with diverse genome structures (4, 6, 7, 16, 19, 20, 23, 25, 40, 41, 44, 51, 54, 60). These studies have demonstrated the importance of cellular chaperones in multiple steps of the viral life cycle, including entry, viral protein translation, genome replication, encapsidation, and virion release. However, the list of virus-hsp connections is likely incomplete. Further studies to explore this particular host-pathogen interaction will shed light on virus replication mechanisms and pathogenesis, and potentially highlight targets for novel antiviral agents.To study the role of cellular chaperones in the genome replication of positive-sense RNA viruses, we use flock house virus (FHV), a natural insect pathogen and well-studied member of the Nodaviridae family. The FHV life cycle shares many common features with other positive-sense RNA viruses, including the membrane-specific targeting and assembly of functional RNA replication complexes (37, 38), the exploitation of various cellular processes and host factors for viral replication (5, 23, 60), and the induction of large-scale membrane rearrangements (24, 28, 38, 39). FHV virions contain a copackaged bipartite genome consisting of RNA1 (3.1 kb) and RNA2 (1.4 kb), which encode protein A, the viral RNA-dependent RNA polymerase, and the structural capsid protein precursor, respectively (1). During active genome replication, FHV produces a subgenomic RNA3 (0.4 kb), which encodes the RNA interference inhibitor protein B2 (12, 29, 32). These viral characteristics make FHV an excellent model system to study many aspects of positive-sense RNA virus biology.In addition to the benefits of a simple genome, FHV is able to establish robust RNA replication in a wide variety of genetically tractable eukaryotic hosts, including Drosophila melanogaster (38), Caenorhabditis elegans (32), and Saccharomyces cerevisiae (46). The budding yeast S. cerevisiae has been an exceptionally useful model host to study the mechanisms of viral RNA replication complex assembly and function with FHV (31, 37, 39, 45, 53, 55, 56, 60) as well as other positive-sense RNA viruses (11). The facile genetics of S. cerevisiae, along with the vast array of well-defined cellular and molecular tools and techniques, make it an ideal eukaryotic host for the identification of cellular factors required for positive-sense RNA virus replication. Furthermore, readily available yeast libraries with deletions and regulated expression of individual proteins have led to the completion of several high-throughput screens to provide a global survey of host factors that impact virus replication (26, 42, 52). An alternative approach with these yeast libraries that reduces the inherently high false-negative rates associated with high-throughput screens is to focus on a select set of host genes associated with a particular cellular pathway, process, or location previously implicated in virus replication.We have utilized such a targeted approach and focused on examining the impact of cytosolic chaperones on FHV RNA replication. Previously, we have shown that the cellular chaperone hsp90 facilitates protein A synthesis in Drosophila cells (5, 23), and the hsp40 encoded by the yeast YDJ1 gene facilitates FHV RNA replication in yeast, in part through effects on both protein A accumulation and function (60). In this report, we further extend these observations by examining FHV RNA accumulation in a panel of yeast strains with deletions of known or hypothesized cytosolic chaperones. We demonstrate that cytosolic chaperones can have either suppressive or enhancing effects on FHV RNA accumulation. In particular, related hsp70 members encoded by the SSA and SSB yeast chaperone families have marked and dramatically divergent effects on both genomic and subgenomic RNA accumulation and viral polymerase synthesis. These results highlight the complexities of the host-pathogen interactions that influence positive-sense RNA virus replication and identify the hsp70 family of cytosolic chaperones as key regulators of FHV replication.  相似文献   

6.
7.
8.
9.
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   

10.
Saccharomyces cerevisiae produces extracellular glycerophosphoinositol through phospholipase-mediated turnover of phosphatidylinositol and transports glycerophosphoinositol into the cell upon nutrient limitation. A screening identified the RAS GTPase-activating proteins Ira1 and Ira2 as required for utilization of glycerophosphoinositol as the sole phosphate source, but the RAS/cyclic AMP pathway does not appear to be involved in the growth phenotype. Ira1 and Ira2 affect both the production and transport of glycerophosphoinositol.Membrane phospholipids are continually synthesized and degraded as cells grow and respond to environmental conditions. A major pathway of phosphatidylinositol (PI) turnover in Saccharomyces cerevisiae is its deacylation to produce extracellular glycerophosphoinositol (GroPIns) (3). Plb3, an enzyme with phospholipase B (PLB)/lysophospholipase activity, is thought to be primarily responsible for the production of extracellular GroPIns, with Plb1 playing a lesser role (11, 12, 13). GroPIns is transported into the cell by the Git1 permease (17). GIT1 expression is upregulated by phosphate limitation and inositol limitation. In fact, GroPIns can act as the cell''s sole source of both inositol (17) and phosphate (1).A screening for gene products involved in the process by which GroPIns enters the cellular metabolism identified Ira1 and Ira2, yeast homologs of the mammalian protein neurofibromin. Alterations in NF1, the gene encoding neurofibromin, are associated with the pathogenesis of neurofibromatosis type 1, an autosomal dominant genetic disease (4, 5, 25). Ira1 and Ira2 and neurofibromin function as RAS GTPase-activating proteins (RAS GAPs). S. cerevisiae Ras1 and Ras2 activate adenylate cyclase to modulate cyclic AMP (cAMP) levels. The binding of cAMP to the regulatory subunits of protein kinase A (Bcy1) results in dissociation and activation of the catalytic subunits (Tpk1 to Tpk3). Ira1 and Ira2 inactivate RAS and thereby downregulate the pathway (18, 19). Hydrolysis of cAMP by the phosphodiesterases encoded by PDE1 and PDE2 also downregulate the pathway (7, 20, 23). The RAS/cAMP pathway responds to nutrient signals to modulate fundamental cellular processes, including stress resistance, metabolism, and cell proliferation (7, 20, 21).  相似文献   

11.
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   

12.
13.
A segregationally stable expression and secretion vector for Saccharomyces cerevisiae, named pYABD01, was constructed by cloning the yeast gene region encoding the mating pheromone α-factor 1 secretion signal (MFα1s) into the S. cerevisiae high-copy-number expression vector pYES2. The structural genes of the two leaderless peptides of enterocin L50 (EntL50A and EntL50B) from Enterococcus faecium L50 were cloned, separately (entL50A or entL50B) and together (entL50AB), into pYABD01 under the control of the galactose-inducible promoter PGAL1. The generation of recombinant S. cerevisiae strains heterologously expressing and secreting biologically active EntL50A and EntL50B demonstrates the suitability of the MFα1s-containing vector pYABD01 to direct processing and secretion of these antimicrobial peptides through the S. cerevisiae Sec system.Lactic acid bacteria (LAB) are widely known for their ability to produce a variety of ribosomally synthesized proteins or peptides, referred to as bacteriocins, displaying antimicrobial activity against a broad range of gram-positive bacteria and, to a lesser extent, gram-negative bacteria, including spoilage and food-borne pathogenic microorganisms (11, 19, 33, 34, 36, 37). These antimicrobials may be classified into three main classes: (i) the lantibiotics, or posttranslationally modified peptides; (ii) the nonmodified, small, heat-stable peptides; and (iii) the large, heat-labile protein bacteriocins. Class II bacteriocins are further grouped into five subclasses: the subclass IIa (pediocin-like bacteriocins containing the N-terminal conserved motif YGNGVxC), the subclass IIb (two-peptide bacteriocins), the subclass IIc (leaderless bacteriocins), the subclass IId (circular bacteriocins), and the subclass IIe (other peptide bacteriocins) (17, 19, 21, 37). All lantibiotics and most class II bacteriocins are synthesized as biologically inactive precursors containing an N-terminal extension (the so-called double-glycine-type leader sequence or the Sec-dependent signal peptide), which is cleaved off concomitantly with externalization of biologically active bacteriocins by a dedicated ATP-binding cassette transporter and its accessory protein or by the Sec system and the signal peptidases, respectively (11, 17). Interestingly, only a few bacteriocins described to date are synthesized without an N-terminal extension, including enterocin L50 (L50A and L50B) (8), enterocin Q (EntQ) (10), enterocin EJ97 (41), and the bacteriocin LsbB (20).In recent years, there has been an increasing interest in the application of bacteriocinogenic microorganisms and/or their bacteriocins as biopreservatives to guarantee the safety and quality of foods and beverages, such as fermented vegetables and meats, dairy and fish products, and wine and beer (12, 15, 16, 39, 47). Three main strategies for the use of bacteriocins as food biopreservatives have been proposed: (i) addition of a purified/semipurified bacteriocin preparation as a food additive; (ii) use of a substrate previously fermented by a bacteriocin-producing strain as a food ingredient; and/or (iii) inoculation of a culture to produce the bacteriocin in situ in fermented foods (13, 15). The lantibiotic nisin A is the most widely characterized bacteriocin and the only one that has been legally approved in more than 48 countries as a food additive for use in certain types of cheeses (13, 16). Likewise, nisin A has been approved as a beer additive in Australia and New Zealand (16). However, the difficulties encountered in addressing the regulatory approval of new bacteriocins as food additives have spurred the development of the other bacteriocin-based food biopreservation strategies (13, 17).Beer is a beverage with a remarkable microbiological stability and is considered as a food substrate difficult to spoil. However, some LAB, such as Lactobacillus brevis, Lactobacillus lindneri, and Pediococcus damnosus, are able to spoil beer and are recognized as the most hazardous bacteria for breweries, being responsible for approximately 70% of microbial beer spoilage incidents (40, 47). The ever-growing consumer demand for less-processed and less chemically preserved foods and beverages is promoting the development of alternative biocontrol strategies, such as those based on the use of bacteriocins as biopreservatives (12, 15, 39, 47). However, beyond the strict requirements to fulfill legal regulations, the commercial application of bacteriocins as beer additives is hindered mainly by low bacteriocin production yields and increases in production costs (44). Considering that Saccharomyces cerevisiae is commonly used as starter culture for brewing (24, 28, 35), a novel beer biopreservation strategy based on the development of bactericidal S. cerevisiae brewing strains has been proposed to overcome the aforementioned challenges (44, 46, 47). In this respect, the heterologous production of LAB bacteriocins, namely, pediocin PA-1 (PedPA-1) from Pediococcus acidilactici PAC1.0 and plantaricin 423 from Lactobacillus plantarum 423, by laboratory strains of S. cerevisiae has been reported (44, 46).Enterocin L50 (EntL50) is a commonly found bacteriocin composed of two highly related leaderless antimicrobial peptides, enterocin L50A (EntL50A) and enterocin L50B (EntL50B), which possesses a broad antimicrobial spectrum against LAB, food-borne pathogenic bacteria, and human and animal clinical pathogens (8, 9, 10, 11). Previous work by our group showed that EntL50 (EntL50A and EntL50B) may be used as a beer biopreservative to inhibit the growth of beer spoilage bacteria (1). Therefore, genetically engineered strains of S. cerevisiae heterologously expressing and secreting EntL50A and EntL50B have been developed in this work. For this purpose, we constructed the segregationally stable expression and secretion vector pYABD01, which allowed the secretion of biologically active EntL50A and EntL50B directed by MFα1s through the S. cerevisiae Sec system.  相似文献   

14.
15.
Staphylococcus aureus is a highly versatile and evolving bacterium of great clinical importance. S. aureus can evolve by acquiring single nucleotide polymorphisms and mobile genetic elements and by recombination events. Identification and location of novel genomic elements in a bacterial genome are not straightforward, unless the whole genome is sequenced. Optical mapping is a new tool that creates a high-resolution, in situ ordered restriction map of a bacterial genome. These maps can be used to determine genomic organization and perform comparative genomics to identify genomic rearrangements, such as insertions, deletions, duplications, and inversions, compared to an in silico (virtual) restriction map of a known genome sequence. Using this technology, we report here the identification, approximate location, and characterization of a genetic inversion of ∼500 kb of a DNA element between the NRS387 (USA800) and FPR3757 (USA300) strains. The presence of the inversion and location of its junction sites were confirmed by site-specific PCR and sequencing. At both the left and right junction sites in NRS387, an IS1181 element and a 73-bp sequence were identified as inverted repeats, which could explain the possible mechanism of the inversion event.Staphylococcus aureus is a gram-positive bacterium of immense clinical importance. This opportunistic pathogen is capable of causing a wide range of diseases from skin and soft-tissue infections to life-threatening bacteremia, endocarditis, and osteomyelitis (14). Approximately 75% of the S. aureus genome is composed of a core genome that is common in all strains, and 25% of the genome is composed of variable regions which can differ between different strains (4, 16, 24-26). S. aureus evolves primarily by introducing single nucleotide polymorphisms in its core genome and by acquiring mobile genetic elements (MGEs) through horizontal gene transfer. These MGEs include pathogenicity/genomic islands, plasmids, transposons, and bacteriophages that become integrated in the chromosome (4, 11, 16, 31, 32). Despite being a heterogeneous organism, genetic recombination in S. aureus is proposed to be rather rare (20, 24, 29, 35). Its clones are more likely to evolve by point mutations than by recombination events (12). The MGEs contribute to the phenotypic and genotypic diversity seen with the S. aureus population. Acquisition of the staphylococcal cassette chromosome (SCCmec) elements through site-specific recombinases has led to the epidemic of methicillin-resistant S. aureus (MRSA) strains in hospitals and communities all over the world (6, 10, 15). In recent years, the integration of arginine catabolite mobile element in the USA300 lineage of MRSA has been proposed to give the pathogen its epidemiological advantage, including traits for surviving in low-pH conditions and oxygen tension environments (11). In addition, chromosomal replacements have been observed within lineages of sequence type 34 (ST34) and ST42 (34) and ST8 and ST30 (13).Genomic rearrangements, such as inversions, have been observed with genomes of enteric bacteria, such as Salmonella enterica, Shigella flexneri, Yersinia pestis KIM, Escherichia coli (K12 and O157:H7), and group A Streptococcus pyogenes (8, 9, 18, 27, 28, 30, 37). No genomic inversions in S. aureus have been reported to date. With the use of optical mapping, large genomic rearrangements, such as inversions, that would otherwise be missed with other comparative genotyping approaches, including microarray analysis, can be identified. Optical mapping uses high-resolution restriction maps (optical maps) of a bacterial genome to determine its genomic organization (5, 21, 23, 33, 36). These optical maps can be compared to an in silico (virtual) restriction map of a known genome sequence and can be used to identify gene rearrangements and their locations. Using optical mapping in conjunction with subsequent site-specific PCR and sequencing, we report the identification, approximate location, and partial characterization of an ∼500-kb DNA element in NRS387, a USA800 strain that was found to be inverted relative to USA300FPR3757. Identification of IS1181 elements and a novel 73-bp element at both ends of the ∼500-kb element in NRS387 could suggest the possibility of an inversion event in an ancestral strain of NRS387.  相似文献   

16.
17.
In their vertebrate hosts, arboviruses such as Semliki Forest virus (SFV) (Togaviridae) generally counteract innate defenses and trigger cell death. In contrast, in mosquito cells, following an early phase of efficient virus production, a persistent infection with low levels of virus production is established. Whether arboviruses counteract RNA interference (RNAi), which provides an important antiviral defense system in mosquitoes, is an important question. Here we show that in Aedes albopictus-derived mosquito cells, SFV cannot prevent the establishment of an antiviral RNAi response or prevent the spread of protective antiviral double-stranded RNA/small interfering RNA (siRNA) from cell to cell, which can inhibit the replication of incoming virus. The expression of tombusvirus siRNA-binding protein p19 by SFV strongly enhanced virus spread between cultured cells rather than virus replication in initially infected cells. Our results indicate that the spread of the RNAi signal contributes to limiting virus dissemination.In animals, RNA interference (RNAi) was first described for Caenorhabditis elegans (27). The production or introduction of double-stranded RNA (dsRNA) in cells leads to the degradation of mRNAs containing homologous sequences by sequence-specific cleavage of mRNAs. Central to RNAi is the production of 21- to 26-nucleotide small interfering RNAs (siRNAs) from dsRNA and the assembly of an RNA-induced silencing complex (RISC), followed by the degradation of the target mRNA (23, 84). RNAi is a known antiviral strategy of plants (3, 53) and insects (21, 39, 51). Study of Drosophila melanogaster in particular has given important insights into RNAi responses against pathogenic viruses and viral RNAi inhibitors (31, 54, 83, 86, 91). RNAi is well characterized for Drosophila, and orthologs of antiviral RNAi genes have been found in Aedes and Culex spp. (13, 63).Arboviruses, or arthropod-borne viruses, are RNA viruses mainly of the families Bunyaviridae, Flaviviridae, and Togaviridae. The genus Alphavirus within the family Togaviridae contains several mosquito-borne pathogens: arboviruses such as Chikungunya virus (16) and equine encephalitis viruses (88). Replication of the prototype Sindbis virus and Semliki Forest virus (SFV) is well understood (44, 71, 74, 79). Their genome consists of a positive-stranded RNA with a 5′ cap and a 3′ poly(A) tail. The 5′ two-thirds encodes the nonstructural polyprotein P1234, which is cleaved into four replicase proteins, nsP1 to nsP4 (47, 58, 60). The structural polyprotein is encoded in the 3′ one-third of the genome and cleaved into capsid and glycoproteins after translation from a subgenomic mRNA (79). Cytoplasmic replication complexes are associated with cellular membranes (71). Viruses mature by budding at the plasma membrane (35).In nature, arboviruses are spread by arthropod vectors (predominantly mosquitoes, ticks, flies, and midges) to vertebrate hosts (87). Little is known about how arthropod cells react to arbovirus infection. In mosquito cell cultures, an acute phase with efficient virus production is generally followed by the establishment of a persistent infection with low levels of virus production (9). This is fundamentally different from the cytolytic events following arbovirus interactions with mammalian cells and pathogenic insect viruses with insect cells. Alphaviruses encode host response antagonists for mammalian cells (2, 7, 34, 38).RNAi has been described for mosquitoes (56) and, when induced before infection, antagonizes arboviruses and their replicons (1, 4, 14, 15, 29, 30, 32, 42, 64, 65). RNAi is also functional in various mosquito cell lines (1, 8, 43, 49, 52). In the absence of RNAi, alphavirus and flavivirus replication and/or dissemination is enhanced in both mosquitoes and Drosophila (14, 17, 31, 45, 72). RNAi inhibitors weakly enhance SFV replicon replication in tick and mosquito cells (5, 33), posing the questions of how, when, and where RNAi interferes with alphavirus infection in mosquito cells.Here we use an A. albopictus-derived mosquito cell line to study RNAi responses to SFV. Using reporter-based assays, we demonstrate that SFV cannot avoid or efficiently inhibit the establishment of an RNAi response. We also demonstrate that the RNAi signal can spread between mosquito cells. SFV cannot inhibit cell-to-cell spread of the RNAi signal, and spread of the virus-induced RNAi signal (dsRNA/siRNA) can inhibit the replication of incoming SFV in neighboring cells. Furthermore, we show that SFV expression of a siRNA-binding protein increases levels of virus replication mainly by enhancing virus spread between cells rather than replication in initially infected cells. Taken together, these findings suggest a novel mechanism, cell-to-cell spread of antiviral dsRNA/siRNA, by which RNAi limits SFV dissemination in mosquito cells.  相似文献   

18.
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   

19.
Bocavirus is a newly classified genus of the family Parvovirinae. Infection with Bocavirus minute virus of canines (MVC) produces a strong cytopathic effect in permissive Walter Reed/3873D (WRD) canine cells. We have systematically characterized the MVC infection-produced cytopathic effect in WRD cells, namely, the cell death and cell cycle arrest, and carefully examined how MVC infection induces the cytopathic effect. We found that MVC infection induces an apoptotic cell death characterized by Bax translocalization to the mitochondrial outer membrane, disruption of the mitochondrial outer membrane potential, and caspase activation. Moreover, we observed that the activation of caspases occurred only when the MVC genome was replicating, suggesting that replication of the MVC genome induces apoptosis. MVC infection also induced a gradual cell cycle arrest from the S phase in early infection to the G2/M phase at a later stage, which was confirmed by the upregulation of cyclin B1 and phosphorylation of cdc2. Cell cycle arrest at the G2/M phase was reproduced by transfection of a nonreplicative NS1 knockout mutant of the MVC infectious clone, as well as by inoculation of UV-irradiated MVC. In contrast with other parvoviruses, only expression of the MVC proteins by transfection did not induce apoptosis or cell cycle arrest. Taken together, our results demonstrate that MVC infection induces a mitochondrion-mediated apoptosis that is dependent on the replication of the viral genome, and the MVC genome per se is able to arrest the cell cycle at the G2/M phase. Our results may shed light on the molecular pathogenesis of Bocavirus infection in general.The Bocavirus genus is newly classified within the subfamily Parvovirinae of the family Parvoviridae (21). The currently known members of the Bocavirus genus include bovine parvovirus type 1 (BPV1) (17), minute virus of canines (MVC) (57), and the recently identified human bocaviruses (HBoV, HBoV2, and HBoV3) (4, 7, 36).MVC was first recovered from canine fecal samples in 1970 (10). The virus causes respiratory disease with breathing difficulty (14, 32, 49) and enteritis with severe diarrhea (11, 39), which often occurs with coinfection with other viruses (39), spontaneous abortion of fetuses, and death of newborn puppies (14, 29). Pathological lesions in fetuses in experimental infections were found in the lymphoid tissue of the lung and small intestine (14). MVC was isolated and grown in the Walter Reed/3873D (WRD) canine cell line (10), which is derived from a subdermoid cyst of an irradiated male dog (10). The full-length 5.4-kb genome of MVC was recently mapped with palindromic termini (60). Under the control of a single P6 promoter, through the mechanism of alternative splicing and alternative polyadenylation, MVC expresses two nonstructural proteins (NS1 and NP1) and two capsid proteins (VP1 and VP2). Like the NS1 proteins of other parvoviruses, the NS1 of MVC is indispensable for genome replication. The NP1 protein, which is unique to the Bocavirus genus, appears to be critical for optimal viral replication, as the NP1 knockout mutant of MVC suffers from severe impairment of replication (60). A severe cytopathic effect during MVC infection of WRD cells has been documented (10, 60).The HBoV genome has been frequently detected worldwide in respiratory specimens from children under 2 years old with acute respiratory illnesses (2, 34, 55). HBoV is associated with acute expiratory wheezing and pneumonia (3, 34, 55) and is commonly detected in association with other respiratory viruses (34, 55). Further studies are necessary, however, to identify potential associations of HBoV infection with clinical symptoms or disease of acute gastroenteritis (7, 36). The full-length sequence of infectious MVC DNA (GenBank accession no. FJ214110) that we have reported shows 52.6% identity to HBoV, while the NS1, NP1, and VP1 proteins are 38.5%, 39.9%, and 43.7% identical to those of HBoV, respectively (60).The cytopathic effect induced during parvovirus infection has been widely documented, e.g., in infections with minute virus of mice (MVM) (13), human parvovirus B19 (B19V) (58), parvovirus H-1 (25, 52), and BPV1 (1). In Bocavirus, cell death during BPV1 infection of embryonic bovine tracheal cells has been shown to be achieved through necrosis, independent of apoptosis (1). B19V-induced cell death of primary erythroid progenitor cells has been shown to be mainly mediated by an apoptotic pathway (58) in which the nonstructural protein 11kDa plays a key role (16). In contrast, the MVM-induced cytopathic effect has been revealed to be mediated by NS1 interference with intracellular casein kinase II (CKII) signaling (22, 44, 45), a nonapoptotic cell death. Oncolytic parvovirus H-1 infections can induce either apoptosis or nonapoptotic cell death, depending on the cell type (25, 40). Therefore, the mechanisms underlying parvovirus infection-induced cell death vary, although NS1 has been widely shown to be involved in both apoptotic and nonapoptotic cell death. The nature of the cytopathic effect during Bocavirus MVC infection has not been studied.Parvovirus replication requires infected cells at the S phase. Infection with parvovirus has been revealed to accompany a cell cycle perturbation that mostly leads to an arrest in the S/G2 phase or the G2/M phase during infection (30, 33, 42, 47, 65). MVM NS1 expression induces an accumulation of sensitive cells in the S/G2 phase (6, 46, 47). Whether MVC infection-induced cell death is accompanied by an alternation of cell cycle progression and whether the viral nonstructural protein is involved in these processes have not been addressed.In this study, we found, in contrast with other members of the family Parvoviridae, expression of both the nonstructural and structural proteins of MVC by transfection did not induce cell death or cell cycle arrest. However, the cytopathic effect induced during MVC infection is a replication-coupled, mitochondrion-mediated and caspase-dependent apoptosis, accompanied with a gradual cell cycle arrest from the S phase to the G2/M phase, which is facilitated by the MVC genome.  相似文献   

20.
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   

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