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 15 毫秒
1.
The 3.0 A resolution crystal structure of Pariacoto virus (PaV) reveals extensive interactions between portions of the viral RNA genome and the icosahedral capsid. Under the protein shell of the T = 3 quasi equivalent capsid lies a dodecahedral cage composed of RNA duplex that accounts for approximately 35% of the single-stranded RNA genome. The highly basic N-terminal regions (residues 7-54) of the subunits, forming pentamers (A subunits) are clearly visible in the density map and make numerous interactions with the RNA cage. The C-terminal segments (residues 394-401) of the A subunits lie in channels near the quasi three-fold axes. Electron cryo-microscopy and image reconstruction of PaV particles clearly show the dodecahedral RNA cage.  相似文献   

2.
Pariacoto virus (PaV) is a nodavirus that was recently isolated in Peru from the Southern armyworm, Spodoptera eridania. Virus particles are non enveloped and about 30 nm in diameter and have T=3 icosahedral symmetry. The 3.0-A crystal structure shows that about 35% of the genomic RNA is icosahedrally ordered, with the RNA forming a dodecahedral cage of 25-nucleotide (nt) duplexes that underlie the inner surface of the capsid. The PaV genome comprises two single-stranded, positive-sense RNAs: RNA1 (3,011 nt), which encodes the 108-kDa catalytic subunit of the RNA-dependent RNA polymerase, and RNA2 (1,311 nt), which encodes the 43-kDa capsid protein precursor alpha. In order to apply molecular genetics to the structure and assembly of PaV, we identified susceptible cell lines and developed a reverse genetic system for this virus. Cell lines that were susceptible to infection by PaV included those from Spodoptera exigua, Helicoverpa zea and Aedes albopictus, whereas cells from Drosophila melanogaster and Spodoptera frugiperda were refractory to infection. To recover virus from molecular clones, full-length cDNAs of PaV RNAs 1 and 2 were cotranscribed by T7 RNA polymerase in baby hamster kidney cells that expressed T7 RNA polymerase. Lysates of these cells were infectious both for cultured cells from Helicoverpa zea (corn earworm) and for larvae of Galleria mellonella (greater wax moth). The combination of infectious cDNA clones, cell culture infectivity, and the ability to produce milligram amounts of virus allows the application of DNA-based genetic methods to the study of PaV structure and assembly.  相似文献   

3.
The nodavirus Flock house virus (FHV) has a bipartite, positive-sense RNA genome that is packaged into an icosahedral particle displaying T=3 symmetry. The high-resolution X-ray structure of FHV has shown that 10 bp of well-ordered, double-stranded RNA are located at each of the 30 twofold axes of the virion, but it is not known which portions of the genome form these duplex regions. The regular distribution of double-stranded RNA in the interior of the virus particle indicates that large regions of the encapsidated genome are engaged in secondary structure interactions. Moreover, the RNA is restricted to a topology that is unlikely to exist during translation or replication. We used electron cryomicroscopy and image reconstruction to determine the structure of four types of FHV particles that differed in RNA and protein content. RNA-capsid interactions were primarily mediated via the N and C termini, which are essential for RNA recognition and particle assembly. A substantial fraction of the packaged nucleic acid, either viral or heterologous, was organized as a dodecahedral cage of duplex RNA. The similarity in tertiary structure suggests that RNA folding is independent of sequence and length. Computational modeling indicated that RNA duplex formation involves both short-range and long-range interactions. We propose that the capsid protein is able to exploit the plasticity of the RNA secondary structures, capturing those that are compatible with the geometry of the dodecahedral cage.  相似文献   

4.
The vesivirus feline calicivirus (FCV) is a positive strand RNA virus encapsidated by an icosahedral T=3 shell formed by the viral VP1 protein. Upon its expression in the insect cell - baculovirus system in the context of vaccine development, two types of virus-like particles (VLPs) were formed, a majority built of 60 subunits (T=1) and a minority probably built of 180 subunits (T=3). The structure of the small particles was determined by x-ray crystallography at 0.8 nm resolution helped by cryo-electron microscopy in order to understand their formation. Cubic crystals belonged to space group P213. Their self-rotation function showed the presence of an octahedral pseudo-symmetry similar to the one described previously by Agerbandje and co-workers for human parvovirus VLPs. The crystal structure could be solved starting from the published VP1 structure in the context of the T=3 viral capsid. In contrast to viral capsids, where the capsomers are interlocked by the exchange of the N-terminal arm (NTA) domain, this domain is disordered in the T=1 capsid of the VLPs. Furthermore it is prone to proteolytic cleavage. The relative orientation of P (protrusion) and S (shell) domains is alerted so as to fit VP1 to the smaller T=1 particle whereas the intermolecular contacts around 2-fold, 3-fold and 5-fold axes are conserved. By consequence the surface of the VLP is very similar compared to the viral capsid and suggests a similar antigenicity. The knowledge of the structure of the VLPs will help to improve their stability, in respect to a use for vaccination.  相似文献   

5.
An assembly intermediate of a small, non-enveloped RNA virus has been discovered that exhibits striking differences from the mature virion. Virus-like particles (VLPs) of Nudaurelia capensis omega virus (NomegaV), a T=4 icosahedral virus infecting Lepidoptera insects, were produced in insect cells using a baculovirus vector expressing the coat protein. A procapsid form was discovered when NomegaV VLPs were purified at neutral pH conditions. These VLPs were fragile and did not undergo the autoproteolytic maturation that occurs in the infectious virus. Electron cryo-microscopy (cryoEM) and image analysis showed that, compared with the native virion, the VLPs were 16% larger in diameter, more rounded, porous, and contained an additional internal domain. Upon lowering the pH to 5.0, the VLP capsids became structurally indistinguishable from the authentic virion and the subunits autoproteolyzed. The NomegaV protein subunit coordinates, which were previously determined crystallographically, were modelled into the 28 A resolution cryoEM map of the procapsid. The resulting pseudo-atomic model of the NomegaV procapsid demonstrated the large rearrangements in quaternary and tertiary structure needed for the maturation of the VLPs and presumably of the virus. Based on this model, we propose that electrostatically driven rearrangements of interior helical regions are responsible for the large conformational change. These results are surprising because large structural rearrangements have not been found in the maturation of any other small RNA viruses. However, similarities of this conformational change to the maturational processes of more complex DNA viruses (e.g. bacteriophages and herpesvirus) and to the swelling of simple plant viruses suggest that structural changes in icosahedral viruses, which are integral to their function, have similar strategies and perhaps mechanisms.  相似文献   

6.
7.
Using cryo-electron microscopy, single particle image processing and three-dimensional reconstruction with icosahedral averaging, we have determined the three-dimensional solution structure of bacteriophage MS2 capsids reassembled from recombinant protein in the presence of short oligonucleotides. We have also significantly extended the resolution of the previously reported structure of the wild-type MS2 virion. The structures of recombinant MS2 capsids reveal clear density for bound RNA beneath the coat protein binding sites on the inner surface of the T = 3 MS2 capsid, and show that a short extension of the minimal assembly initiation sequence that promotes an increase in the efficiency of assembly, interacts with the protein capsid forming a network of bound RNA. The structure of the wild-type MS2 virion at ∼9 Å resolution reveals icosahedrally ordered density encompassing ∼90% of the single-stranded RNA genome. The genome in the wild-type virion is arranged as two concentric shells of density, connected along the 5-fold symmetry axes of the particle. This novel RNA fold provides new constraints for models of viral assembly.  相似文献   

8.
Infectious bursal disease virus (IBDV) is a nonenveloped avian virus with a two-segment double-stranded RNA genome. Its T=13 icosahedral capsid is most probably assembled with 780 subunits of VP2 and 600 copies of VP3 and has a diameter of about 60 nm. VP1, the RNA-dependent RNA polymerase, resides inside the viral particle. Using a baculovirus expression system, we first observed that expression of the pVP2-VP4-VP3 polyprotein encoded by the genomic segment IBDA results mainly in the formation of tubules with a diameter of about 50 nm and composed of pVP2, the precursor of VP2. Very few virus-like particles (VLPs) and VP4 tubules with a diameter of about 25 nm were also identified. The inefficiency of VLP assembly was further investigated by expression of additional IBDA-derived constructs. Expression of pVP2 without any other polyprotein components results in the formation of isometric particles with a diameter of about 30 nm. VLPs were observed mainly when a large exogeneous polypeptide sequence (the green fluorescent protein sequence) was fused to the VP3 C-terminal domain. Large numbers of VLPs were visualized by electron microscopy, and single particles were shown to be fluorescent by standard and confocal microscopy analysis. Moreover, the final maturation process converting pVP2 into the VP2 mature form was observed on generated VLPs. We therefore conclude that the correct scaffolding of the VP3 can be artificially induced to promote the formation of VLPs and that the final processing of pVP2 to VP2 is controlled by this particular assembly. To our knowledge, this is the first report of the engineering of a morphogenesis switch to control a particular type of capsid protein assembly.  相似文献   

9.
10.
Flock house virus (FHV) is a bipartite, positive-strand RNA insect virus that encapsidates its two genomic RNAs in a single virion. It provides a convenient model system for studying the principles underlying the copackaging of multipartite viral RNA genomes. In this study, we used a baculovirus expression system to determine if the uncoupling of viral protein synthesis from RNA replication affected the packaging of FHV RNAs. We found that neither RNA1 (which encodes the viral replicase) nor RNA2 (which encodes the capsid protein) were packaged efficiently when capsid protein was supplied in trans from nonreplicating RNA. However, capsid protein synthesized in cis from replicating RNA2 packaged RNA2 efficiently in the presence and absence of RNA1. These results demonstrated that capsid protein translation from replicating RNA2 is required for specific packaging of the FHV genome. This type of coupling between genome replication and translation and RNA packaging has not been observed previously. We hypothesize that RNA2 replication and translation must be spatially coordinated in FHV-infected cells to facilitate retrieval of the viral RNAs for encapsidation by newly synthesized capsid protein. Spatial coordination of RNA and capsid protein synthesis may be key to specific genome packaging and assembly in other RNA viruses.  相似文献   

11.
Killer toxin-secreting strains of the yeasts Hanseniaspora uvarum and Zygosaccharomyces bailii were shown to contain linear double-stranded RNAs (dsRNAs) that persist within the cytoplasm of the infected host cell as encapsidated virus-like particles. In both yeasts, L- and M-dsRNAs were associated with 85-kDa major capsid protein, whereas the additional Z-dsRNA (2.8 kb), present only in the wild-type Z. bailii killer strain, was capsid protein, whereas the additional Z-dsRNA (2.8 kb), present only in the wild-type Z. bailii killer strain, was shown to be encapsidated by a 35-kDa coat protein. Although Northern (RNA) blot hybridizations indicated that L-dsRNA from Z. bailii is a LA species, additional peptide maps of the purified 85-kDa capsid from Z. bailii and the 88- and 80-kDa major coat proteins from K1 and K28 killer viruses of Saccharomyces cerevisiae revealed distinctly different patterns of peptides. Electron microscopy of purified Z. bailii viruses (ZbV) identified icosahedral particles 40 nm in diameter which were undistinguishable from the S. cerevisiae killer viruses. We demonstrated that purified ZbVs are sufficient to confer the Z. bailii killer phenotype on transfected spheroplasts of a S. cerevisiae nonkiller strain and that the resulting transfectants secreted even more killer toxin that the original ZbV donor strain did. Curing experiments with ZbV-transfected S. cerevisiae strains indicated that the M-dsRNA satellite from Z. bailii contains the genetic information for toxin production, whereas expression of toxin immunity might be dependent on Z-dsRNA, which resembles a new dsRNA replicon in yeasts that is not dependent on an LA helper virus to be stably maintained and replicated within the cell.  相似文献   

12.
In Sindbis virus, initiation of nucleocapsid core assembly begins with recognition of the encapsidation signal of the viral RNA genome by capsid protein. This nucleation event drives the recruitment of additional capsid proteins to fully encapsidate the genome, generating an icosahedral nucleocapsid core. The encapsidation signal of the Sindbis virus genomic RNA has previously been localized to a 132-nucleotide region of the genome within the coding region of the nsP1 protein, and the RNA-binding activity of the capsid was previously mapped to a central region of the capsid protein. It is unknown how capsid protein binding to encapsidation signal leads to ordered oligomerization of capsid protein and nucleocapsid core assembly. To address this question, we have developed a mobility shift assay to study this interaction. We have characterized a 32 amino acid peptide capable of recognizing the Sindbis virus encapsidation signal RNA. Using this peptide, we were able to observe a conformational change in the RNA induced by capsid protein binding. Binding is tight (K(d)(app) = 12 nM), and results in dimerization of the capsid peptide. Mutational analysis reveals that although almost every predicted secondary structure within the encapsidation signal is required for efficient protein binding, the identities of the bases within the helices and hairpin turns of the RNA do not need to be maintained. In contrast, two purine-rich loops are essential for binding. From these data, we have developed a model in which the encapsidation signal RNA adopts a highly folded structure and this folding process directs early events in nucleocapsid assembly.  相似文献   

13.
In this work, we evaluate the stability, dynamics and protein-nucleic acid interaction in Flock House virus (FHV). FHV is an RNA insect virus, non-enveloped, member of the family Nodaviridae. It is composed of a bipartite single-stranded RNA genome packaged in an icosahedral capsid of 180 copies of an identical protein (alpha protein). A fundamental property of many animal viruses is the post-assembly maturation required for infectivity. FHV is constructed as a provirion, which matures to an infectious virion by cleavage of alpha protein into beta and gamma subunits. We used high pressure, temperature and chemical denaturing agents to promote perturbation of the viral capsid. These effects were monitored by spectroscopy measurements (fluorescence, light scattering and CD) and size-exclusion chromatography. The data showed that FHV was stable to pressures up to 310 MPa at room temperature. The fluorescence emission and light scattering values showed small changes that were reversible after decompression. When we combined pressure and sub-denaturing urea concentrations (1 M), the changes were more drastic, suggesting dissociation of the capsid. However, these changes were reversible after pressure release. The complete dissociation of FHV could be observed only under high urea concentrations (10 M). There were no significant changes in emission spectra up to 5 M urea. FHV also was stable when we used temperature treatments (high and low). We also compared the effects of urea and pressure on FHV wild type and cleavage-defective mutant VLPs (virus-like particles). The VLPs and authentic particles are distinguishable by protein-RNA interactions, since VLPs pack cellular RNA and native particles contain viral RNA. Our results demonstrated that native particles are more stable than VLPs to physical and chemical treatments. Our data point to the specificity of the interaction between the capsid protein and the viral RNA. This specificity is crucial to the stability of the particle, which makes this interaction an excellent target for drug development.  相似文献   

14.
Macrobrachium rosenbergii nodavirus (MrNV) is a virus native to giant freshwater prawn. Recombinant MrNV capsid protein has been produced in Escherichia coli, which self‐assembled into virus‐like particles (VLPs). However, this recombinant protein is unstable, degrading and forming heterogenous VLPs. In this study, MrNV capsid protein was produced in insect Spodoptera frugiperda (Sf9) cells through a baculovirus system. Dynamic light scattering (DLS) and transmission electron microscopy (TEM) revealed that the recombinant protein produced by the insect cells self‐assembled into highly stable, homogenous VLPs each of approximately 40 nm in diameter. Enzyme‐linked immunosorbent assay (ELISA) showed that the VLPs produced in Sf9 cells were highly antigenic and comparable to those produced in E. coli. In addition, the Sf9 produced VLPs were highly stable across a wide pH range (2–12). Interestingly, the Sf9 produced VLPs contained DNA of approximately 48 kilo base pairs and RNA molecules. This study is the first report on the production and characterization of MrNV VLPs produced in a eukaryotic system. The MrNV VLPs produced in Sf9 cells were about 10 nm bigger and had a uniform morphology compared with the VLPs produced in E. coli. The insect cell production system provides a good source of MrNV VLPs for structural and immunological studies as well as for host–pathogen interaction studies. © 2016 American Institute of Chemical Engineers Biotechnol. Prog., 33:549–557, 2017  相似文献   

15.
将兔出血症病毒衣壳蛋白VP6 0基因插入杆状病毒转移载体pBLUEBACHIS2_B的 6 HIS表达标签下游 ,与线性化野生型杆状病毒基因组DNA共转染Sf9昆虫细胞 ,经蚀斑纯化后获克隆化重组杆状病毒pBLUEBACHIS2B_VP6 0。以重组杆状病毒感染Sf9细胞 ,经SDS_PAGE和Westernblot检测显示高效表达一分子量为 6 9kD的重组蛋白 ,并且该蛋白可被兔抗RHDV高免血清识别。血凝试验表明 ,该重组蛋白可以凝集人“O”型红细胞 ,血凝价达 2 1 6 ,同时 ,该血凝性可被抗RHDV的高免血清所抑制。经电镜观察 ,重组病毒表达的融合有 6 HIS表达标签的衣壳蛋白仍可在昆虫细胞内自聚成不包裹核酸的、与天然RHDV病毒粒子在物理形态上相似的病毒样颗粒 (VLPs) ,并且该VLPs与兔抗RHDV高免血清作用后于电镜下可见凝集成团的现象 ,表明其与天然RHDV病毒粒子在抗原性上也极为相似  相似文献   

16.
Defective interfering (DI) RNA genomes of poliovirus which contain in-frame deletions in the P1 capsid protein-encoding region have been described. DI genomes are capable of replication and can be encapsidated by capsid proteins provided in trans from wild-type poliovirus. In this report, we demonstrate that a previously described poliovirus DI genome (K. Hagino-Yamagishi and A. Nomoto, J. Virol. 63:5386-5392, 1989) can be complemented by a recombinant vaccinia virus, VVP1 (D. C. Ansardi, D. C. Porter, and C. D. Morrow, J. Virol. 65:2088-2092, 1991), which expresses the poliovirus capsid precursor polyprotein, P1. Stocks of defective polioviruses were generated by transfecting in vitro-transcribed defective genome RNA derived from plasmid pSM1(T7)1 into HeLa cells infected with VVP1 and were maintained by serial passage in the presence of VVP1. Encapsidation of the defective poliovirus genome was demonstrated by characterizing poliovirus-specific protein expression in cells infected with preparations of defective poliovirus and by Northern (RNA) blot analysis of poliovirus-specific RNA incorporated into defective poliovirus particles. Cells infected with preparations of defective poliovirus expressed poliovirus protein 3CD but did not express capsid proteins derived from a full-length P1 precursor. Poliovirus-specific RNA encapsidated in viral particles generated in cells coinfected with VVP1 and defective poliovirus migrated slightly faster on formaldehyde-agarose gels than wild-type poliovirus RNA, demonstrating maintenance of the genomic deletion. By metabolic radiolabeling with [35S]methionine-cysteine, the defective poliovirus particles were shown to contain appropriate mature-virion proteins. This is the first report of the generation of a pure population of defective polioviruses free of contaminating wild-type poliovirus. We demonstrate the use of this recombinant vaccinia virus-defective poliovirus genome complementation system for studying the effects of a defined mutation in the P1 capsid precursor on virus assembly. Following removal of residual VVP1 from defective poliovirus preparations, processing and assembly of poliovirus capsid proteins derived from a nonmyristylated P1 precursor expressed by a recombinant vaccinia virus, VVP1 myr- (D. C. Ansardi, D. C. Porter, and C. D. Morrow, J. Virol. 66:4556-4563, 1992), in cells coinfected with defective poliovirus were analyzed. Capsid proteins generated from nonmyristylated P1 did not assemble detectable levels of mature virions but did assemble, at low levels, into empty capsids.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

17.
18.
The secondary structure of encapsidated MS2 genomic RNA poses an interesting RNA folding challenge. Cryoelectron microscopy has demonstrated that encapsidated MS2 RNA is well-ordered. Models of MS2 assembly suggest that the RNA hairpin-protein interactions and the appropriate placement of hairpins in the MS2 RNA secondary structure can guide the formation of the correct icosahedral particle. The RNA hairpin motif that is recognized by the MS2 capsid protein dimers, however, is energetically unfavorable, and thus free energy predictions are biased against this motif. Computer programs called Crumple, Sliding Windows, and Assembly provide useful tools for prediction of viral RNA secondary structures when the traditional assumptions of RNA structure prediction by free energy minimization may not apply. These methods allow incorporation of global features of the RNA fold and motifs that are difficult to include directly in minimum free energy predictions. For example, with MS2 RNA the experimental data from SELEX experiments, crystallography, and theoretical calculations of the path for the series of hairpins can be incorporated in the RNA structure prediction, and thus the influence of free energy considerations can be modulated. This approach thoroughly explores conformational space and generates an ensemble of secondary structures. The predictions from this new approach can test hypotheses and models of viral assembly and guide construction of complete three-dimensional models of virus particles.  相似文献   

19.
Formation of an RNA Heterodimer upon Heating of Nodavirus Particles   总被引:5,自引:3,他引:2       下载免费PDF全文
Flock House virus is a small icosahedral insect virus of the family Nodaviridae. Its genome consists of two positive-sense RNA molecules, which are believed to be encapsidated into a single viral particle. However, evidence to support this claim is circumstantial. Here we demonstrate that exposure of nodavirus particles to heat causes the two strands of viral RNA to form a stable complex, directly establishing that both RNAs are copackaged into one virion. The physical properties of the RNA complex, the effect of heat on the particles per se, and the possible relevance of these findings to the nodavirus life cycle are presented.  相似文献   

20.
To investigate the role of protein-protein and protein-nucleic acid interactions in virus assembly, we compared the stabilities of native bacteriophage MS2, virus-like particles (VLPs) containing nonviral RNAs, and an assembly-defective coat protein mutant (dlFG) and its single-chain variant (sc-dlFG). Physical (high pressure) and chemical (urea and guanidine hydrochloride) agents were used to promote virus disassembly and protein denaturation, and the changes in virus and protein structure were monitored by measuring tryptophan intrinsic fluorescence, bis-ANS probe fluorescence, and light scattering. We found that VLPs dissociate into capsid proteins that remain folded and more stable than the proteins dissociated from authentic particles. The proposed model is that the capsid disassembles but the protein remains bound to the heterologous RNA encased by VLPs. The dlFG dimerizes correctly, but fails to assemble into capsids, because it lacks the 15-amino acid FG loop involved in inter-dimer interactions at the viral fivefold and quasi-sixfold axes. This protein was very unstable and, when compared with the dissociation/denaturation of the VLPs and the wild-type virus, it was much more susceptible to chemical and physical perturbation. Genetic fusion of the two subunits of the dimer in the single-chain dimer sc-dlFG stabilized the protein, as did the presence of 34-bp poly(GC) DNA. These studies reveal mechanisms by which interactions in the capsid lattice can be sufficiently stable and specific to ensure assembly, and they shed light on the processes that lead to the formation of infectious viral particles.  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