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.
2.
Norovirus GII/4 is a leading cause of acute viral gastroenteritis in humans. We examined here how the GII/4 virus evolves to generate and sustain new epidemics in humans, using 199 near-full-length GII/4 genome sequences and 11 genome segment clones from human stool specimens collected at 19 sites in Japan between May 2006 and February 2009. Phylogenetic studies demonstrated outbreaks of 7 monophyletic GII/4 subtypes, among which a single subtype, termed 2006b, had continually predominated. Phylogenetic-tree, bootscanning-plot, and informative-site analyses revealed that 4 of the 7 GII/4 subtypes were mosaics of recently prevalent GII/4 subtypes and 1 was made up of the GII/4 and GII/12 genotypes. Notably, single putative recombination breakpoints with the highest statistical significance were constantly located around the border of open reading frame 1 (ORF1) and ORF2 (P ≤ 0.000001), suggesting outgrowth of specific recombinant viruses in the outbreaks. The GII/4 subtypes had many unique amino acids at the time of their outbreaks, especially in the N-term, 3A-like, and capsid proteins. Unique amino acids in the capsids were preferentially positioned on the outer surface loops of the protruding P2 domain and more abundant in the dominant subtypes. These findings suggest that intersubtype genome recombination at the ORF1/2 boundary region is a common mechanism that realizes independent and concurrent changes on the virion surface and in viral replication proteins for the persistence of norovirus GII/4 in human populations.Norovirus (NoV) is a nonenveloped RNA virus that belongs to the family Caliciviridae and can cause acute gastroenteritis in humans. The NoV genome is a single-stranded, positive-sense, polyadenylated RNA that encodes three open reading frames, ORF1, ORF2, and ORF3 (68). ORF1 encodes a long polypeptide (∼200 kDa) that is cleaved in the cells by the viral proteinase (3Cpro) into six proteins (4). These proteins function in NoV replication in host cells (19). ORF2 encodes a viral capsid protein, VP1. The capsid gene evolved at a rate of 4.3 × 10−3 nucleotide substitutions/site/year (7), which is comparable to the substitution rates of the envelope and capsid genes of human immunodeficiency virus (30). The capsid protein of NoV consists of a shell (S) and two protruding (P) domains: P1 and P2 (47). The S domain is relatively conserved within the same genetic lineages of NoVs (38) and is responsible for the assembly of VP1 (6). The P1 subdomain is also relatively conserved (38) and has a role in enhancing the stability of virus particles (6). The P2 domain is positioned at the most exposed surface of the virus particle (47) and forms binding clefts for putative infection receptors, such as human histo-blood group antigens (HBGA) (8, 13, 14, 60). The P2 domain also contains epitopes for neutralizing antibodies (27, 33) and is consistently highly variable even within the same genetic lineage of NoVs (38). ORF3 encodes a VP2 protein that is suggested to be a minor structural component of virus particles (18) and to be responsible for the expression and stabilization of VP1 (5).Thus far, the NoVs found in nature are classified into five genogroups (GI to GV) and multiple genotypes on the basis of the phylogeny of capsid sequences (71). Among them, genogroup II genotype 4 (GII/4), which was present in humans in the mid-1970s (7), is now the leading cause of NoV-associated acute gastroenteritis in humans (54). The GII/4 is further subclassifiable into phylogenetically distinct subtypes (32, 38, 53). Notably, the emergence and spread of a new GII/4 subtype with multiple amino acid substitutions on the capsid surface are often associated with greater magnitudes of NoV epidemics (53, 54). In 2006 and 2007, a GII/4 subtype, termed 2006b, prevailed globally over preexisting GII/4 subtypes in association with increased numbers of nonbacterial acute gastroenteritis cases in many countries, including Japan (32, 38, 53). The 2006b subtype has multiple unique amino acid substitutions that occur most preferentially in the protruding subdomain of the capsid, the P2 subdomain (32, 38, 53). Together with information on human population immunity against NoV GII/4 subtypes (12, 32), it has been postulated that the accumulation of P2 mutations gives rise to antigenic drift and plays a key role in new epidemics of NoV GII/4 in humans (32, 38, 53).Genetic recombination is common in RNA viruses (67). In NoV, recombination was first suggested by the phylogenetic analysis of an NoV genome segment clone: a discordant branching order was noted with the trees of the 3Dpol and capsid coding regions (21). Subsequently, many studies have reported the phylogenetic discordance using sequences from various epidemic sites in different study periods (1, 10, 11, 16, 17, 22, 25, 40, 41, 44-46, 49, 51, 57, 63, 64, 66). These results suggest that genome recombination frequently occurs among distinct lineages of NoV variants in vivo. However, the studies were done primarily with direct sequencing data of the short genome portion, and information on the cloned genome segment or full-length genome sequences is very limited (21, 25). Therefore, we lack an overview of the structural and temporal dynamics of viral genomes during NoV epidemics, and it remains unclear whether NoV mosaicism plays a role in these events.To clarify these issues, we collected 199 near-full-length genome sequences of GII/4 from NoV outbreaks over three recent years in Japan, divided them into monophyletic subtypes, analyzed the temporal and geographical distribution of the subtypes, collected phylogenetic evidence for the viral genome mosaicism of the subtypes, identified putative recombination breakpoints in the genomes, and isolated mosaic genome segments from the stool specimens. We also performed computer-assisted sequence and structural analyses with the identified subtypes to address the relationship between the numbers of P2 domain mutations at the times of the outbreaks and the magnitudes of the epidemics. The obtained data suggest that intersubtype genome recombination at the ORF1/2 boundary region is common in the new GII/4 outbreaks and promotes the effective acquisition of mutation sets of heterogeneous capsid surface and viral replication proteins.  相似文献   

3.
We have sequenced the genome and identified the structural proteins and lipids of the novel membrane-containing, icosahedral virus P23-77 of Thermus thermophilus. P23-77 has an ∼17-kb circular double-stranded DNA genome, which was annotated to contain 37 putative genes. Virions were subjected to dissociation analysis, and five protein species were shown to associate with the internal viral membrane, while three were constituents of the protein capsid. Analysis of the bacteriophage genome revealed it to be evolutionarily related to another Thermus phage (IN93), archaeal Halobacterium plasmid (pHH205), a genetic element integrated into Haloarcula genome (designated here as IHP for integrated Haloarcula provirus), and the Haloarcula virus SH1. These genetic elements share two major capsid proteins and a putative packaging ATPase. The ATPase is similar with the ATPases found in the PRD1-type viruses, thus providing an evolutionary link to these viruses and furthering our knowledge on the origin of viruses.Three-dimensional structures of the major capsid proteins, as well as the architecture of the virion and the sequence similarity of putative genome packaging ATPases, have revealed unexpected evolutionary connection between virus families. Viruses infecting hosts residing in different domains of life (Bacteria, Archaea, and Eukarya) share common structural elements and possibly also ways to package the viral genome (8, 13, 41). It has been proposed that the set of genes responsible for virion assembly is a hallmark of the virus and is designated as the innate viral “self,” which may retain its identity through evolutionary times (5). Based on this, it is proposed that viruses can be classified into lineages that span the different domains of life. Therefore, the studies of new virus isolates might provide insights into the events that led to the origin of viruses and maybe even the origin of life itself (34, 40). However, viruses are known to be genetic mosaics (28), and these structural lineages therefore do not reflect the evolutionary history of all genes in a given virus. For example, the genome replication strategies vary significantly even in the currently established lineages (41) and, consequently, a structural approach does not point out to a specific form of replication in the ancestor. Nevertheless, as the proposal for a viral self is driven from information on viral structures and pathways of genome encapsidation, the ancestral form of the self was likely to be composed of a protective coat and the necessary mechanisms to incorporate the genetic material within the coat.Viruses structurally related to bacteriophage PRD1, a phage infecting gram-negative bacteria, have been identified in all three domains of life, and the lineage hypothesis was first proposed based on structural information on such viruses. Initially, PRD1 and human adenovirus were proposed to originate from a common ancestor mainly due to the same capsid organization (T=25) and the major coat protein topology, the trimeric double β-barrel fold (12). In addition, these viruses share a common vertex organization and replication mechanism (20, 31, 53, 63). PRD1 is an icosahedral virus with an inner membrane, whereas adenovirus lacks the membrane. Later, many viruses with similar double β-barrel fold in the major coat protein have been discovered and included to this viral lineage. For example, the fold is present in Paramecium bursaria Chlorella virus 1 (56) of algae, Bam35 (45) of gram-positive bacteria, PM2 (2) of gram-negative marine bacteria, and Sulfolobus turreted icosahedral virus (STIV) (38) of an archaeal host. Moreover, genomic analyses have revealed a common set of genes in a number of nucleocytoplasmic large DNA viruses. Chilo iridescent virus and African swine fever virus 1 are related to Paramecium bursaria Chlorella virus 1 and most probably share structural similarity to PRD1-type viruses (13, 30, 31, 68). The largest known viruses, represented by mimivirus and poxvirus, may also belong to this lineage (29, 77). Two euryarchaeal proviruses, TKV4 and MVV, are also proposed to belong to this lineage based on bioinformatic searches (42). The proposed PRD1-related viruses share the same basic architectural principles despite major differences in the host organisms and particle and genome sizes (1, 2, 38, 56). PM2, for example, has a genome of only 10 kbp, whereas mimivirus (infecting Acanthamoeba polyphaga) double-stranded DNA (dsDNA) genome is 1.2 Mbp in size (59).How many virion structure-based lineages might there be? This obviously relates to the number of protein folds that have the properties needed to make viral capsids. It has been noted that, in addition to PRD1-type viruses, at least tailed bacterial and archaeal viruses, as well as herpesviruses, share the same coat protein fold. Also, certain dsRNA viruses seem to have structural and functional similarities, although their hosts include bacteria and yeasts, as well as plants and animals (6, 18, 19, 27, 55, 60, 74). Obviously, many structural principles to build a virus capsid exist, and it has been suggested that especially geothermally heated environments have preserved many of the anciently formed virus morphotypes (35).Thermophilic dsDNA bacteriophage P23-77 was isolated from an alkaline hot spring in New Zealand on Thermus thermophilus (17) ATCC 33923 (deposited as Thermus flavus). P23-77 was shown to have an icosahedral capsid and possibly an internal membrane but no tail (81). Previously, another Thermus virus, IN93, with a similar morphology has been described (50). IN93 was inducible from a lysogenic strain of Thermus aquaticus TZ2, which was isolated from hot spring soil in Japan. Recently, P23-77 was characterized in more detail (33). It has an icosahedral protein coat, organized in a T=28 capsid lattice (21). The presence of an internal membrane was confirmed, and lipids were shown to be constituents of the virion. Ten structural proteins were identified, with apparent molecular masses ranging from 8 to 35 kDa. Two major protein species with molecular masses of 20 and 35 kDa were proposed to make the capsomers, one forming the hexagonal building blocks and the other the two towers that decorate the capsomer bases (33). Surprisingly, P23-77 is structurally closest to the haloarchaeal virus SH1, which is the only other example of a T=28 virion architecture (32, 33). In both cases it was proposed that the capsomers are made of six single β-barrels opposing the situation with the other structurally related viruses where the hexagonal capsomers are made of three double β-barrel coat protein monomers (8).In the present study we analyze the dsDNA genome of P23-77. Viral membrane proteins and those associated with the capsid were identified by virion dissociation studies. The protein chemistry data and genome annotation are consistent with the results of the disruption studies. A detailed analysis of the lipid composition of P23-77 and its T. thermophilus host was carried out. The data collected here reveal additional challenges in attempts to generate viral lineages based on the structural and architectural properties of the virion.  相似文献   

4.
5.
Like other RNA viruses, coxsackievirus B5 (CVB5) exists as circulating heterogeneous populations of genetic variants. In this study, we present the reconstruction and characterization of a probable ancestral virion of CVB5. Phylogenetic analyses based on capsid protein-encoding regions (the VP1 gene of 41 clinical isolates and the entire P1 region of eight clinical isolates) of CVB5 revealed two major cocirculating lineages. Ancestral capsid sequences were inferred from sequences of these contemporary CVB5 isolates by using maximum likelihood methods. By using Bayesian phylodynamic analysis, the inferred VP1 ancestral sequence dated back to 1854 (1807 to 1898). In order to study the properties of the putative ancestral capsid, the entire ancestral P1 sequence was synthesized de novo and inserted into the replicative backbone of an infectious CVB5 cDNA clone. Characterization of the recombinant virus in cell culture showed that fully functional infectious virus particles were assembled and that these viruses displayed properties similar to those of modern isolates in terms of receptor preferences, plaque phenotypes, growth characteristics, and cell tropism. This is the first report describing the resurrection and characterization of a picornavirus with a putative ancestral capsid. Our approach, including a phylogenetics-based reconstruction of viral predecessors, could serve as a starting point for experimental studies of viral evolution and might also provide an alternative strategy for the development of vaccines.The group B coxsackieviruses (CVBs) (serotypes 1 to 6) were discovered in the 1950s in a search for new poliovirus-like viruses (33, 61). Infections caused by CVBs are often asymptomatic but may occasionally result in severe diseases of the heart, pancreas, and central nervous system (99). CVBs are small icosahedral RNA viruses belonging to the Human enterovirus B (HEV-B) species within the family Picornaviridae (89). In the positive single-stranded RNA genome, the capsid proteins VP1 to VP4 are encoded within the P1 region, whereas the nonstructural proteins required for virus replication are encoded within the P2 and P3 regions (4). The 30-nm capsid has an icosahedral symmetry and consists of 60 copies of each of the four structural proteins. The VP1, VP2, and VP3 proteins are surface exposed, whereas the VP4 protein lines the interior of the virus capsid (82). The coxsackievirus and adenovirus receptor (CAR), a cell adhesion molecule of the immunoglobulin superfamily, serves as the major cell surface attachment molecule for all six serotypes of CVB (5, 6, 39, 60, 98). Some strains of CVB1, CVB3 and CVB5 also interact with the decay-accelerating factor (DAF) (CD55), a member of the family of proteins that regulate the complement cascade. However, the attachment of CVBs to DAF alone does not permit the infection of cells (6, 7, 59, 85).Picornaviruses exist as genetically highly diverse populations within their hosts, referred to as quasispecies (20, 57). This genetic plasticity enables these viruses to adapt rapidly to new environments, but at the same time, it may compromise the structural integrity and enzymatic functionality of the virus. The selective constraints imposed on the picornavirus genome are reflected in the different regions used for different types of evolutionary studies. The highly conserved RNA-dependent RNA polymerase (3Dpol) gene is used to establish phylogenetic relationships between more-distantly related viruses (e.g., viruses belonging to different genera) (38), whereas the variable genomic sequence encoding the VP1 protein is used for the classification of serotypes (13, 14, 69, 71, 72).In 1963, Pauling and Zuckerkandl proposed that comparative analyses of contemporary protein sequences can be used to predict the sequences of their ancient predecessors (73). Experimental reconstruction of ancestral character states has been applied to evolutionary studies of several different proteins, e.g., galectins (49), G protein-coupled receptors (52), alcohol dehydrogenases (95), rhodopsins (15), ribonucleases (46, 88, 110), elongation factors (32), steroid receptors (10, 96, 97), and transposons (1, 45, 87). In the field of virology, reconstructed ancestral or consensus protein sequences have been used in attempts to develop vaccine candidates for human immunodeficiency virus type 1 (21, 51, 66, 81) but rarely to examine general phenotypic properties.In this study, a CVB5 virus with a probable ancestral virion (CVB5-P1anc) was constructed and characterized. We first analyzed in detail the evolutionary relationships between structural genes of modern CVB5 isolates and inferred a time scale for their evolutionary history. An ancestral virion sequence was subsequently inferred by using a maximum likelihood (ML) method. This sequence was then synthesized de novo, cloned into a replicative backbone of an infectious CVB5 cDNA clone, and transfected into HeLa cells. The hypothetical CVB5-P1anc assembled into functional virus particles that displayed phenotypic properties similar to those of contemporary clinical isolates. This is the first report describing the reconstruction and characterization of a fully functional picornavirus with a putative ancestral capsid.  相似文献   

6.
7.
8.
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   

9.
10.
After recognizing and binding to its host cell, poliovirus (like other nonenveloped viruses) faces the challenge of translocating its genome across a cellular membrane and into the cytoplasm. To avoid entanglement with the capsid, the RNA must exit via a single site on the virion surface. However, the mechanism by which a single site is selected (from among 60 equivalents) is unknown; and until now, even its location on the virion surface has been controversial. To help to elucidate the mechanism of infection, we have used single-particle cryo-electron microscopy and tomography to reconstruct conformationally altered intermediates that are formed by the poliovirion at various stages of the poliovirus infection process. Recently, we reported icosahedrally symmetric structures for two forms of the end-state 80S empty capsid particle. Surprisingly, RNA was frequently visible near the capsid; and in a subset of the virions, RNA was seen on both the inside and outside of the capsid, caught in the act of exiting. To visualize RNA exiting, we have now determined asymmetric reconstructions from that subset, using both single-particle cryo-electron microscopy and cryo-electron tomographic methods, producing independent reconstructions at ∼50-Å resolution. Contrary to predictions in the literature, the footprint of RNA on the capsid surface is located close to a viral 2-fold axis, covering a slot-shaped area of reduced density that is present in both of the symmetrized 80S reconstructions and which extends by about 20 Å away from the 2-fold axis toward each neighboring 5-fold axis.In its role as the intermediate that links one round of infection with the next, a virus particle protects the viral genome during passage from cell to cell and from host to host, it specifically recognizes and binds to target cells, and it delivers the viral genome into the appropriate compartment in the target cell. For enveloped viruses, which have their own external membranes, fusion of the viral membrane with a host membrane presents a conceptually simple mechanism for delivery of the genome or nucleoprotein into the cytoplasm. For nonenveloped viruses, the viral particle must provide the machinery necessary for either the entire virion, a nucleoprotein complex, or the viral genome to cross a membrane. This process remains poorly understood. Poliovirus provides an excellent model system for probing the mechanisms used for genome translocation. As the type member of the Picornavirus family and the etiological agent of poliomyelitis, poliovirus has been well characterized biochemically and genetically (42), its cell entry pathways have been well characterized (5, 15, 30, 52), and a number of cell entry intermediates have been identified and are accessible for structural studies (2-4, 7, 8, 18, 34, 38, 42, 55, 56).The capsid of the mature poliovirion (160S particle) consists of 60 copies of each of the four coat proteins VP1, VP2, VP3, and VP4 (which is myristolated at its amino-terminal glycine [13]) and encloses a 7.5-kbp positive-sense RNA genome. The outer surface of the capsid has a number of major features, including star-shaped mesas at its 5-fold axes, 3-fold propeller-like protrusions, canyon-like depressions surrounding each of the 5-fold mesas, and depressions at the 2-fold axes (30, 31).Poliovirus infection is initiated when the virus binds to the host-cell-surface poliovirus receptor (called Pvr or CD155) (41), triggering a conformational change of the native capsid into an altered particle called the A particle or 135S particle (18, 19). The 135S particle has been shown to be expanded by about 4% (2, 7), is infectious (16, 33), and is believed to be a productive intermediate in viral entry (30, 33). This conformational change results in the externalization of the small myristoylated capsid protein, VP4 (18), and of the amino-terminal extension of VP1 (which includes a conserved amphipathic helix) (23). Both of these externalized polypeptides then associate with membranes (17, 23). In subsequent steps, the viral genome is released from the capsid and translocated across a membrane (probably an endosomal membrane [5]) to gain access to the cytoplasm, leaving behind an end-state empty capsid shell (called the 80S particle). The trigger for RNA release and the mechanism of genome translocation are both poorly understood (30, 52).Electrophysiology and mutational experiments have shown that the externalization of VP4 and of the amino terminus of VP1 is associated with the formation of channels in membranes (17, 49, 50) and, furthermore, that point mutations in threonine 28 of VP4 can either eliminate (T28G) or alter (T28V, T28S) the ability to form channels and either eliminate (T28G) or slow (T28V, T28S) the kinetics of productive RNA release (17). These observations have led to the hypothesis that the viral polypeptides insert into host cell membranes during infection and rearrange to form channels that permit the viral genome to pass through the membrane, thereby gaining access to the cytoplasm (7, 17, 49, 50).Speculation about the sites of externalization of the viral peptides and of the viral genome began soon after the structures of mature rhinovirus and poliovirus were determined crystallographically 25 years ago (31, 44). In both structures there is a solvent-filled channel running along each 5-fold axis. This channel is closed off at the outer surface of the capsid by polypeptide loops and on the inner surface by a plug that is formed by five intertwined copies of the amino terminus of VP3, forming a parallel beta tube (31, 44). In poliovirus this tube is flanked on its inner surface by five copies of a three-stranded beta sheet in which the outermost two strands come from a beta hairpin at the amino terminus of VP4 and the innermost strand comes from residues at the extreme amino terminus of VP1 (20). The presence of this channel, together with its proximity to peptide segments that were known to be externalized upon receptor attachment, and analogies with other viruses led to a model in which both the peptides and the viral RNA are externalized via the channel at the 5-fold axis (25, 45). At that time, an alternative model for the egress of polypeptides was proposed, based on an analogy with the externalization of the amino-terminal extensions of capsid proteins in expanded states of the topologically similar T=3 plant viruses (26, 32, 43, 47) and on genetic and biochemical studies of mutations that affect cell entry and capsid stability in poliovirus (14, 39, 54). In the latter model, the peptides were proposed to exit from the base of the canyon and then proceed along the outer surface toward the 5-fold peak (43, 47). Both models suggested that five copies of each of the externalized peptides would interact in some way to form a pore in the membrane that was contiguous with one of the 5-fold channels, thus providing a way for RNA to be released from the virion at a 5-fold axis of symmetry. No data yet exist to specify what specific structural roles VP4 and the amino terminus of VP1 might play in forming pores and serving as membrane anchors. However, both the electrophysiology data (cited above) and the greater sequence conservation of VP4 suggest that its role in pore formation may be the more central (17, 49, 50).To further elucidate various steps along the infection pathway, cryo-electron microscopy (cryo-EM) reconstructions have been determined for a number of cell entry intermediates of poliovirus and rhinoviruses, and their resolutions have been improved over time (2, 3, 7, 28, 38). Structures of the complexes of polioviruses and major-group rhinoviruses with the ectodomains of their respective receptors have confirmed earlier models that suggested that the canyon is the receptor-binding site and have begun to suggest how receptor binding might lead to receptor-induced conformational rearrangements (3, 56). Cryo-EM and cryo-electron tomography structures (cryo-ET) of a poliovirus-receptor-membrane complex (using a novel receptor-decorated liposome model [51]) confirmed that initial receptor binding brings the surface of the 5-fold mesa into close proximity with the membrane and appears to produce an outward distortion of the outer leaflet of the membrane in its area of closest approach to the virus particle (4, 8).Structures have also been determined for the soluble 135S and 80S particles of poliovirus, formed by heating the virus at 50°C (135S) or 56°C (80S) in hypotonic buffers, and for the 80S particles of rhinovirus 14 and 16, formed by exposing virus to acidic pH. All of the biological and immunological evidence that is currently available indicates that the particles prepared in vitro and used for structural studies are indistinguishable from the particles that are released from the cell surface during infection (6, 53). These structures have allowed the models for peptide release and genome release to be extended and refined (7, 38) and indeed have confirmed that VP1 exits from the particle surface at the base of the canyon and climbs up the side of the 5-fold mesa. However, contrary to the assumptions of the earlier models, the 10-Å structures of the poliovirus 135S and 80S particles show that the amino end of the amino-terminal extension of VP1 does not remain associated with the mesa. Instead, it forms an alpha-helical bridge that stretches across the canyon and binds to the large EF loop of VP2, a surface projection that appears as a 3-fold propeller blade (7, 38).Until recently, the mechanism of RNA release (during the 135S-to-80S transition) has been largely a matter of conjecture. We can infer that the RNA must exit via a single site on the virion surface, to avoid entanglement with the capsid (particularly as entanglement has never been observed in electron micrographs), though the mechanism by which a single site is selected (from among 60 equivalents) is unknown. All models presented to date have assumed that the RNA is released from the channel at the 5-fold axes (2, 3, 7, 8, 25, 27, 28, 30, 42, 45). However, in the icosahedrally constrained 10-Å structures of both the poliovirus 135S and 80S particles (7, 38), the apparent intactness and stability of the 5-fold mesa argues against the 5-fold axis being the site of RNA egress, given that the diameter of the opening, as seen in those structures, would be insufficient to accommodate RNA, even if the “plug” formed by the intertwined amino termini of VP3 was displaced. Moreover, both structures revealed significant thinning between 2-fold-related pentamers in the vicinity of the 2-fold axes. Most convincingly, large holes (easily sufficient to accommodate RNA) were seen at and near the 2-fold axes in the atomic model of the late-80S structure. This coincided with an open hole in the reconstruction, when viewed at a contour level that left most of the remainder of the capsid intact. This evidence was suggestive, but not definitive, as a number of other openings were present, particularly in the interfaces between protomers. Furthermore, the behavior of the capsid structure in the immediate vicinity of the unique site of RNA exit is likely to be different from what we see in the icosahedral average, which is dominated by the remainder of the capsid.In the course of solving icosahedrally symmetric cryo-EM structures for the poliovirus end-state 80S empty capsid particle (7, 38), we were surprised to find that RNA was frequently visible near the capsid and that in a subset of about 5% of the sampled virions, RNA was seen on both the inside and outside of the capsid, apparently caught in the act of exiting. This was an exciting development, as images of viral RNA release had never previously been reported. We were able to improve the resolution to ∼10 Å by classifying the projected images into two groups: an early 80Se particle that was more prevalent in the population after a shorter heating time and a late 80Sl particle that was seen more often when the heating time was increased. The amount of RNA density remaining in the interior appears to be continuously variable in both classes, suggesting that release is gradual. Of the 5% subset of particles clearly caught in the act, almost all belonged to the 80Se class. Our interpretation was that the 80Se class may represent particles in which exiting RNA is still engaged with the capsid machinery and traversing the capsid, while the 80Sl class (in which much of the capsid resembles the 135S form more closely in structure) represents particles with the RNA disengaged, possibly after nuclease cleavage. More than two structural classes may be present, but at the current resolution, we could not distinguish them.The present report addresses the question of what we can learn about the details of RNA release from an asymmetric cryo-EM reconstruction, based on the 540-particle caught-in-the-act subset, and independently from cryo-electron tomographic reconstructions of a similarly prepared sample. In each projected particle image or subtomogram, preliminary orientation parameters are first determined from an icosahedrally symmetric calculation, and in a second stage, the symmetry is broken by choosing 1 of the 60 symmetry-equivalent orientations. Both methods have yielded similar information, at about 50-Å resolution, concerning the footprint of the RNA on the virion surface, which demonstrates that RNA is released from an asymmetric site at the base of the canyon near a particle 2-fold axis and not at the channel at the 5-fold axes, as suggested by previous models. Additionally, the demonstrated success of the methodology provides us with a blueprint for resolving the molecular details of the RNA-capsid interaction in future experiments.  相似文献   

11.
12.
13.
14.
The first morphological evidence of African swine fever virus (ASFV) assembly is the appearance of precursor viral membranes, thought to derive from the endoplasmic reticulum, within the assembly sites. We have shown previously that protein p54, a viral structural integral membrane protein, is essential for the generation of the viral precursor membranes. In this report, we study the role of protein p17, an abundant transmembrane protein localized at the viral internal envelope, in these processes. Using an inducible virus for this protein, we show that p17 is essential for virus viability and that its repression blocks the proteolytic processing of polyproteins pp220 and pp62. Electron microscopy analyses demonstrate that when the infection occurs under restrictive conditions, viral morphogenesis is blocked at an early stage, immediately posterior to the formation of the viral precursor membranes, indicating that protein p17 is required to allow their progression toward icosahedral particles. Thus, the absence of this protein leads to an accumulation of these precursors and to the delocalization of the major components of the capsid and core shell domains. The study of ultrathin serial sections from cells infected with BA71V or the inducible virus under permissive conditions revealed the presence of large helicoidal structures from which immature particles are produced, suggesting that these helicoidal structures represent a previously undetected viral intermediate.African swine fever virus (ASFV) (61, 72) is the only known DNA-containing arbovirus and the sole member of the Asfarviridae family (24). Infection by this virus of its natural hosts, the wild swine warthogs and bushpigs and the argasid ticks of the genus Ornithodoros, results in a mild disease, often asymptomatic, with low viremia titers, that in many cases develops into a persistent infection (3, 43, 71). In contrast, infection of domestic pigs leads to a lethal hemorrhagic fever for which the only available methods of disease control are the quarantine of the affected area and the elimination of the infected animals (51).The ASFV genome is a lineal molecule of double-stranded DNA of 170 to 190 kbp in length with convalently closed ends and terminal inverted repeats. The genome encodes more than 150 open reading frames, half of which lack any known or predictable function (16, 75).The virus particle, with an overall icosahedral shape and an average diameter of 200 nm (11), is organized in several concentric layers (6, 11, 15) containing more than 50 structural proteins (29). Intracellular particles are formed by an inner viral core, which contains the central nucleoid surrounded by a thick protein coat, referred to as core shell. This core is enwrapped by an inner lipid envelope (7, 34) on top of which the icosahedral capsid is assembled (26, 27, 31). Extracellular virions possess an additional membrane acquired during the budding from the plasma membrane (11). Both forms of the virus, intracellular and extracellular, are infective (8).The assembly of ASFV particles occurs in the cytoplasm of the infected cell, in viral factories located close to the cell nucleus (6, 13, 49). ASFV factories possess several characteristics similar to those of the cellular aggresomes (35), which are accumulations of aggregates of cellular proteins that form perinuclear inclusions (44).Current models propose that ASFV assembly begins with the modification of endoplasmic reticulum (ER) membranes, which are subsequently recruited to the viral factories and transformed into viral precursor membranes. These ER-derived viral membranes represent the precursors of the inner viral envelope and are the first morphological evidence of viral assembly (7, 60). ASFV viral membrane precursors evolve into icosahedral intermediates and icosahedral particles by the progressive assembly of the outer capsid layer at the convex face of the precursor membranes (5, 26, 27, 31) through an ATP- and calcium-dependent process (19). At the same time, the core shell is formed underneath the concave face of the viral envelope, and the viral DNA and nucleoproteins are packaged and condensed to form the innermost electron-dense nucleoid (6, 9, 12, 69). However, the assembly of the capsid and the internal envelope appears to be largely independent of the components of the core of the particle, since the absence of the viral polyprotein pp220 during assembly produces empty virus-like particles that do not contain the core (9).Comparative genome analysis suggests that ASFV shares a common origin with the members of the proposed nucleocytoplasmic large DNA viruses (NCLDVs) (40, 41). The reconstructed phylogeny of NCLDVs as well as the similitude in the structures and organizations of the genomes indicates that ASFV is more closely related to poxviruses than to other members of the NCLDVs. A consensus about the origin and nature of the envelope of the immature form of vaccinia virus (VV), the prototypical poxvirus, seems to be emerging (10, 17, 20, 54). VV assembly starts with the appearance of crescent-shaped structures within specialized regions of the cytoplasm also known as viral factories (21, 23). The crescent membranes originate from preexisting membranes derived from some specialized compartment of the ER (32, 37, 52, 53, 67), and an operative pathway from the ER to the crescent membrane has recently been described (38, 39). VV crescents apparently grow in length while maintaining the same curvature until they become closed circles, spheres in three dimensions, called immature virions (IV) (22). The uniform curvature is produced by a honeycomb lattice of protein D13L (36, 70), which attaches rapidly to the membranes so that nascent viral membranes always appear to be coated over their entirety. The D13L protein is evolutionarily related to the capsid proteins of the other members of the NCLDV group, including ASFV, but lacks the C-terminal jelly roll motif (40). This structural difference is probably related to the fact that poxviruses are the only member of this group without an icosahedral capsid; instead, the spherical D13L coat acts as a scaffold during the IV stage but is discarded in subsequent steps of morphogenesis (10, 28, 46, 66). Thus, although crescents in VV and precursors of the inner envelope in ASFV are the first morphogenetic stages discernible in the viral factories of these viruses, they seem to be different in nature. Crescents are covered by the D13L protein and are more akin to the icosahedral intermediates of ASFV assembly, whereas ASFV viral membrane precursors are more similar to the naked membranes seen when VV morphogenesis is arrested by rifampin treatment (33, 47, 48, 50) or when the expression of the D13L and A17L proteins are repressed during infection with lethal conditional VV viruses (45, 55, 56, 68, 74, 76).Although available evidence strongly supports the reticular origin of the ASFV inner envelope (7, 60), the mechanism of acquisition remains unknown, and the number of membranes present in the inner envelope is controversial. The traditional view of the inner envelope as formed by two tightly opposed membranes derived from ER collapsed cisternae (7, 59, 60) has recently been challenged by the careful examination of the width of the internal membrane of viral particles and the single outer mitochondrial membrane, carried out using chemical fixation, cryosectioning, and high-pressure freezing (34). The results suggest that the inner envelope of ASFV is a single lipid bilayer, which raises the question of how such a structure can be generated and stabilized in the precursors of the ASFV internal envelope. In the case of VV, the coat of the D13L protein has been suggested to play a key role in the stabilization of the single membrane structure of the crescent (10, 17, 36), but the ASFV capsid protein p72 is not a component of the viral membrane precursors. The identification and functional characterization of the proteins involved in the generation of these structures are essential for the understanding of the mechanisms involved in these early stages of viral assembly. For this reason, we are focusing our interest on the study of abundant structural membrane proteins that reside at the inner envelope of the viral particle. We have shown previously that one of these proteins, p54, is essential for the recruitment of ER membranes to the viral factory (59). Repression of protein p54 expression has a profound impact on virus production and leads to an early arrest in virion morphogenesis, resulting in the virtual absence of membranes in the viral factory.Protein p17, encoded by the late gene D117L in the BA71V strain, is an abundant structural protein (60, 65). Its sequence, which is highly conserved among ASFV isolates (16), does not show any significant similarity with the sequences present in the databases. Protein p17 is an integral membrane protein (18) that is predicted to insert in membranes with a Singer type I topology and has been localized in the envelope precursors as well as in both intracellular and extracellular mature particles (60), suggesting that it resides at the internal envelope, the only membranous structure of the intracellular particles.In this work, we analyze the role of protein p17 in viral assembly by means of an IPTG (isopropyl-β-d-thiogalactopyranoside)-dependent lethal conditional virus. The data presented indicate that protein p17 is essential for viral morphogenesis. The repression of this protein appears to block assembly at the level of viral precursor membranes, resulting in their accumulation at the viral factory.From the electron microscopy analysis of serial sections of viral factories at very early times during morphogenesis, we present experimental evidence that suggests that, during assembly, viral precursor membranes and core material organize into large helicoidal intermediates from which icosahedral particles emerge. The possible role of these structures during ASFV morphogenesis is discussed.  相似文献   

15.
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   

16.
Only a few archaeal viruses have been subjected to detailed structural analyses. Major obstacles have been the extreme conditions such as high salinity or temperature needed for the propagation of these viruses. In addition, unusual morphotypes of many archaeal viruses have made it difficult to obtain further information on virion architectures. We used controlled virion dissociation to reveal the structural organization of Halorubrum pleomorphic virus 1 (HRPV-1) infecting an extremely halophilic archaeal host. The single-stranded DNA genome is enclosed in a pleomorphic membrane vesicle without detected nucleoproteins. VP4, the larger major structural protein of HRPV-1, forms glycosylated spikes on the virion surface and VP3, the smaller major structural protein, resides on the inner surface of the membrane vesicle. Together, these proteins organize the structure of the membrane vesicle. Quantitative lipid comparison of HRPV-1 and its host Halorubrum sp. revealed that HRPV-1 acquires lipids nonselectively from the host cell membrane, which is typical of pleomorphic enveloped viruses.In recent years there has been growing interest in viruses infecting hosts in the domain Archaea (43). Archaeal viruses were discovered 35 years ago (52), and today about 50 such viruses are known (43). They represent highly diverse virion morphotypes in contrast to the vast majority (96%) of head-tail virions among the over 5,000 described bacterial viruses (1). Although archaea are widespread in both moderate and extreme environments (13), viruses have been isolated only for halophiles and anaerobic methanogenes of the kingdom Euryarchaeota and hyperthermophiles of the kingdom Crenarchaeota (43).In addition to soil and marine environments, high viral abundance has also been detected in hypersaline habitats such as salterns (i.e., a multipond system where seawater is evaporated for the production of salt) (19, 37, 50). Archaea are dominant organisms at extreme salinities (36), and about 20 haloarchaeal viruses have been isolated to date (43). The majority of these are head-tail viruses, whereas electron microscopic (EM) studies of highly saline environments indicate that the two other described morphotypes, spindle-shaped and round particles, are the most abundant ones (19, 37, 43). Thus far, the morphological diversity of the isolated haloarchaeal viruses is restricted compared to viruses infecting hyperthermophilic archaea, which are classified into seven viral families (43).All of the previously described archaeal viruses have a double-stranded DNA (dsDNA) genome (44). However, a newly characterized haloarchaeal virus, Halorubrum pleomorphic virus 1 (HRPV-1), has a single-stranded DNA (ssDNA) genome (39). HRPV-1 and its host Halorubrum sp. were isolated from an Italian (Trapani, Sicily) solar saltern. Most of the studied haloarchaeal viruses lyse their host cells, but persistent infections are also typical (40, 44). HRPV-1 is a nonlytic virus that persists in the host cells. In liquid propagation, nonsynchronous infection cycles of HRPV-1 lead to continuous virus production until the growth of the host ceases, resulting in high virus titers in the growth medium (39).The pleomorphic virion of HRPV-1 represents a novel archaeal virus morphotype constituted of lipids and two major structural proteins VP3 (11 kDa) and VP4 (65 kDa). The genome of HRPV-1 is a circular ssDNA molecule (7,048 nucleotides [nt]) containing nine putative open reading frames (ORFs). Three of them are confirmed to encode structural proteins VP3, VP4, and VP8, which is a putative ATPase (39). The ORFs of the HRPV-1 genome show significant similarity, at the amino acid level, to the minimal replicon of plasmid pHK2 of Haloferax sp. (20, 39). Furthermore, an ∼4-kb region, encoding VP4- and VP8-like proteins, is found in the genomes of two haloarchaea, Haloarcula marismortui and Natronomonas pharaonis, and in the linear dsDNA genome (16 kb) of spindle-shaped haloarchaeal virus His2 (39). The possible relationship between ssDNA virus HRPV-1 and dsDNA virus His2 challenges the classification of viruses, which is based on the genome type among other criteria (15, 39).HRPV-1 is proposed to represent a new lineage of pleomorphic enveloped viruses (39). A putative representative of this lineage among bacterial viruses might be L172 of Acholeplasma laidlawii (14). The enveloped virion of L172 is pleomorphic, and the virus has a circular ssDNA genome (14 kb). In addition, the structural protein pattern of L172 with two major structural proteins, of 15 and 53 kDa, resembles that of HRPV-1.The structural approach has made it possible to reveal relationships between viruses where no sequence similarity can be detected. It has been realized that several icosahedral viruses infecting hosts in different domains of life share common virion architectures and folds of their major capsid proteins. These findings have consequences for the concept of the origin of viruses. A viral lineage hypothesis predicts that viruses within the same lineage may have a common ancestor that existed before the separation of the cellular domains of life (3, 5, 8, 26). Currently, limited information is available on the detailed structures of viruses infecting archaea. For example, the virion structures of nontailed icosahedral Sulfolobus turreted icosahedral virus (STIV) and SH1 have been determined (21, 23, 46). However, most archaeal viruses represent unusual, sometimes nonregular, morphotypes (43), which makes it difficult to apply structural methods that are based on averaging techniques.A biochemical approach, i.e., controlled virion dissociation, gives information on the localization and interaction of virion components. In the present study, controlled dissociation was used to address the virion architecture of HRPV-1. A comparative lipid analysis of HRPV-1 and its host was also carried out. Our results show that the unique virion type is composed of a flexible membrane decorated with the glycosylated spikes of VP4 and internal membrane protein VP3. The circular ssDNA genome resides inside the viral membrane vesicle without detected association to any nucleoproteins.  相似文献   

17.
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   

18.
The purpose of the present study was to investigate the inhibition of Vibrio by Roseobacter in a combined liquid-surface system. Exposure of Vibrio anguillarum to surface-attached roseobacters (107 CFU/cm2) resulted in significant reduction or complete killing of the pathogen inoculated at 102 to 104 CFU/ml. The effect was likely associated with the production of tropodithietic acid (TDA), as a TDA-negative mutant did not affect survival or growth of V. anguillarum.Antagonistic interactions among marine bacteria are well documented, and secretion of antagonistic compounds is common among bacteria that colonize particles or surfaces (8, 13, 16, 21, 31). These marine bacteria may be interesting as sources for new antimicrobial drugs or as probiotic bacteria for aquaculture.Aquaculture is a rapidly growing sector, but outbreaks of bacterial diseases are a limiting factor and pose a threat, especially to young fish and invertebrates that cannot be vaccinated. Because regular or prophylactic administration of antibiotics must be avoided, probiotic bacteria are considered an alternative (9, 18, 34, 38, 39, 40). Several microorganisms have been able to reduce bacterial diseases in challenge trials with fish or fish larvae (14, 24, 25, 27, 33, 37, 39, 40). One example is Phaeobacter strain 27-4 (17), which inhibits Vibrio anguillarum and reduces mortality in turbot larvae (27). The antagonism of Phaeobacter 27-4 and the closely related Phaeobacter inhibens is due mainly to the sulfur-containing tropolone derivative tropodithietic acid (TDA) (2, 5), which is also produced by other Phaeobacter strains and Ruegeria mobilis (28). Phaeobacter and Ruegeria strains or their DNA has been commonly found in marine larva-rearing sites (6, 17, 28).Phaeobacter and Ruegeria (Alphaproteobacteria, Roseobacter clade) are efficient surface colonizers (7, 11, 31, 36). They are abundant in coastal and eutrophic zones and are often associated with algae (3, 7, 41). Surface-attached Phaeobacter bacteria may play an important role in determining the species composition of an emerging biofilm, as even low densities of attached Phaeobacter strain SK2.10 bacteria can prevent other marine organisms from colonizing solid surfaces (30, 32).In continuation of the previous research on roseobacters as aquaculture probiotics, the purpose of this study was to determine the antagonistic potential of Phaeobacter and Ruegeria against Vibrio anguillarum in liquid systems that mimic a larva-rearing environment. Since production of TDA in liquid marine broth appears to be highest when roseobacters form an air-liquid biofilm (5), we addressed whether they could be applied as biofilms on solid surfaces.  相似文献   

19.
20.
Penaeus stylirostris densovirus (PstDNV), a pathogen of penaeid shrimp, causes significant damage to farmed and wild shrimp populations. In contrast to other parvoviruses, PstDNV probably has only one type of capsid protein that lacks the phospholipase A2 activity that has been implicated as a requirement during parvoviral host cell infection. The structure of recombinant virus-like particles, composed of 60 copies of the 37.5-kDa coat protein, the smallest parvoviral capsid protein reported thus far, was determined to 2.5-Å resolution by X-ray crystallography. The structure represents the first near-atomic resolution structure within the genus Brevidensovirus. The capsid protein has a β-barrel “jelly roll” motif similar to that found in many icosahedral viruses, including other parvoviruses. The N-terminal portion of the PstDNV coat protein adopts a “domain-swapped” conformation relative to its twofold-related neighbor similar to the insect parvovirus Galleria mellonella densovirus (GmDNV) but in stark contrast to vertebrate parvoviruses. However, most of the surface loops have little structural resemblance to any of the known parvoviral capsid proteins.The Parvoviridae family is a family of small DNA viruses that is divided into two subfamilies, the Parvovirinae that infect vertebrates and the Densovirinae that infect invertebrates. Penaeus stylirostris densovirus (PstDNV), also known as infectious hypodermal and hematopoietic necrosis virus (IHHNV), belongs to the Densovirinae subfamily and was first reported as a highly lethal disease of juvenile shrimp in 1983 (22). The virus has significant commercial impact on the shrimp farming industry, causing mass mortality and severe deformations in penaeid shrimp during catastrophic epidemics in marine aquaculture facilities worldwide (14). PstDNV is closely related to the mosquito brevidensoviruses (35), which have the potential to be used as biological control agents of mosquito-borne diseases, such as malaria (30), dengue, chikungunya, and yellow fever (8).The single-stranded DNA genome of parvoviruses is encapsidated within a nonenveloped, icosahedral protein shell of less than 280 Å in external diameter. The capsid consists of 60 structurally equivalent subunits that are composed of the major viral coat protein and a few copies of N-terminally extended variants of the major capsid protein. A phospholipase A2 (PLA2) activity in the unique N-terminal extension of the largest minor capsid protein plays a crucial role during parvoviral host cell infection (7, 12, 13, 20, 46). The structures of the major capsid protein of several vertebrate parvoviruses have previously been determined to near-atomic resolution (1, 18, 23, 37, 41, 43, 44). However, the only high-resolution structure available for the invertebrate subfamily is that of the insect parvovirus Galleria mellonella densovirus (GmDNV) (36). The central motif of parvoviral capsid proteins is an eight-stranded, antiparallel β-barrel “jelly roll” fold. The surface of the virion, however, is formed by large insertions connecting the strands of the β-barrel, thereby creating features that govern antigenicity, receptor binding, and most intersubunit contacts. Surface characteristics common to most parvoviruses are protrusions at or around the icosahedral threefold axes, depressions on the twofold axes, and canyons surrounding the fivefold axes. At each fivefold apex, a cylindrical pore connects the interior of the virus particle with its exterior surroundings. In full virions, these pores are occupied by a glycine-rich motif in the N-terminal region of the major capsid protein, presumably positioning the N-terminal peptide for externalization. The general surface topology of GmDNV is smoother, probably due to smaller loop insertions. The structure of some of these insertions has diverged from vertebrate parvoviruses beyond recognition (4, 36). The N-terminal portions of twofold-related subunits in GmDNV have swapped their positions relative to those of the vertebrate parvoviruses. A cryo-electron microscopy (cryo-EM) study of Aedes albopictus densovirus, a brevidensovirus, has shown that its surface features are different from GmDNV and the mammalian parvoviruses, in particular in having prominent protrusions at the fivefold axes (9).Although it has been reported that PstDNV contains four structural proteins, as determined by SDS-polyacrylamide gel electrophoresis (3), these data do not fit the coding sequence (35). The 4.1-kb DNA genome of PstDNV (3) encodes in the 3′ half of the plus strand just one structural protein of 329 amino acids, as of now the smallest reported parvoviral capsid protein, and in the 5′ half of the plus strand two nonstructural proteins (666 and 363 amino acids) (35). Having only a single type of capsid protein is an unusual feature for viruses in the Parvoviridae family, where capsids are generally reported to contain two or more coat protein variants. A stretch of 11 amino acids in the N-terminal region of the capsid protein (17-DAHNEDEEHAE-27) is reminiscent of the PLA2 catalytic site (35), but it lacks important conserved motifs of PLA2s. Consequently, but curiously, PstDNV does not have the enzymatic activity that has previously been described as a requirement for parvoviral infectivity.We report here the three-dimensional (3D) crystal structure of recombinant, empty virus-like particles (VLPs) of the shrimp parvovirus PstDNV at 2.5-Å resolution. The loops connecting the strands of the structurally conserved jelly roll motif differ considerably in structure and length from other parvoviruses. The near-atomic resolution structure might provide the basis for the design of capsid binding antiviral compounds that may protect shrimp against parvoviral infection (16, 32, 42). Furthermore, the structure might aid the targeting of monoclonal antibodies to gain functional data about the role of the Brevidensovirus capsid protein during the infection cycle. Such information in turn may permit the design of densovirus-based delivery systems for drugs or pest control agents in aquacultural facilities. The small dimensions of PstDNV VLPs can be advantageous for their possible use as nanoparticles for antigen presentation and transport of immune stimulatory substances or interfering RNAs (21, 26). Additionally, the small size of the PstDNV capsid protein makes the system attractive as a model for studying assembly mechanisms of icosahedral virus capsids.  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