首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
Virus Particles from Conidia of Penicillium Species   总被引:1,自引:0,他引:1       下载免费PDF全文
Virus particles and their component double-stranded ribonucleic acid (dsRNA) have been isolated from conidia and mycelia of certain Penicillium species. The conidia and mycelia of P. stoloniferum NRRL 5267 contained 75 and 85 mug of dsRNA/g (dry weight), respectively. Of the total dsRNA released from NRRL 5267 conidia, 10% was nonencapsulated. Conidia of P. brevi-compactum NRRL 5260 and P. chrysogenum Q-176 contained 2 and 120 mug of dsRNA/g (dry weight), respectively, whereas mycelium from the two species contained 3 and 95 mug of dsRNA/g (dry weight), respectively. No viruses were isolated from conidia or mycelia of P. stoloniferum NRRL 859. A method is described for disruption of both conidia and mycelia. The technique facilitates the isolation and characterization of fungal viruses and their component dsRNA and also potentiates surveying of fungal isolates for the presence of virus.  相似文献   

3.
Double-stranded (ds)RNA fungal viruses are currently assigned to six different families. Those from the family Totiviridae are characterized by nonsegmented genomes and single-layer capsids, 300–450 Å in diameter. Helminthosporium victoriae virus 190S (HvV190S), prototype of recently recognized genus Victorivirus, infects the filamentous fungus Helminthosporium victoriae (telomorph: Cochliobolus victoriae), which is the causal agent of Victoria blight of oats. The HvV190S genome is 5179 bp long and encompasses two large, slightly overlapping open reading frames that encode the coat protein (CP, 772 aa) and the RNA-dependent RNA polymerase (RdRp, 835 aa). To our present knowledge, victoriviruses uniquely express their RdRps via a coupled termination–reinitiation mechanism that differs from the well-characterized Saccharomyces cerevisiae virus L-A (ScV-L-A, prototype of genus Totivirus), in which the RdRp is expressed as a CP/RdRp fusion protein due to ribosomal frameshifting. Here, we used transmission electron cryomicroscopy and three-dimensional image reconstruction to determine the structures of HvV190S virions and two types of virus-like particles (capsids lacking dsRNA and capsids lacking both dsRNA and RdRp) at estimated resolutions of 7.1, 7.5, and 7.6 Å, respectively. The HvV190S capsid is thin and smooth, and contains 120 copies of CP arranged in a “T = 2” icosahedral lattice characteristic of ScV-L-A and other dsRNA viruses. For aid in our interpretations, we developed and used an iterative segmentation procedure to define the boundaries of the two, chemically identical CP subunits in each asymmetric unit. Both subunits have a similar fold, but one that differs from ScV-L-A in many details except for a core α-helical region that is further predicted to be conserved among many other totiviruses. In particular, we predict the structures of other victoriviruses to be highly similar to HvV190S and the structures of most if not all totiviruses including, Leishmania RNA virus 1, to be similar as well.  相似文献   

4.
5.
An experimental procedure developed to concentrate fungal viruses from large volumes of homogenized mycelia with water-soluble polymers, such as polyethylene glycol, possesses advantages over more conventional methods. Concentration of the Penicillium stoloniferum fast-moving virus from mycelial homogenates after addition of polyethylene glycol was rapid and produced large quantities of pure virus.  相似文献   

6.
Qiu L  Li Y  Liu Y  Gao Y  Qi Y  Shen J 《Fungal biology》2010,114(5-6):507-513
Many cultivated mushroom strains, such as Pleurotus ostreatus TD300, displayed symptoms of degeneration. A spherical virus POSV and four dsRNA segments were extracted from mycelium of P. ostreatus TD300. POSV had a diameter of 23 nm and encapsidated a 2.5kb dsRNA segment with coat proteins whose molecular weights were 39 kDa and 30 kDa. Four dsRNA segments were 8.2 kb, 2.5 kb, 2.0 kb, and 1.1 kb in size, respectively. The 1.1 kb dsRNA segment often escaped detection. The cDNA and the amino acid sequences of the 8.2 kb dsRNA were homologous to those of RNA-dependent RNA polymerases (RDRP) of ssRNA oyster mushroom spherical virus (OMSV), and contained conserved motifs A to D which were almost identical to those in RDRP of OMSV. The cDNA and amino acid sequences of the 2.5 kb and 2.0 kb dsRNA segments were homologous to that of RDRP and capsid protein of dsRNA virus P. ostreatus virus 1 (PoV1), respectively. In particular, the amino acid sequence of 2.5 kb dsRNA segment had high identity with the conserved motifs A to C in RDRP of PoV1, a Partiviridae virus. After eliminating the viruses in P. ostreatus TD300, the symptoms of degeneration completely disappeared. The results reveal that P. ostreatus TD300 was at least infected by a particle virus POSV, and two naked viruses, one was a dsRNA virus with a 2.0 kb dsRNA segment, the other was an ssRNA virus whose replicating form of genome was an 8.2 kb dsRNA segment. Mycoviruses infection is a causative agent of mushroom strain degeneration.  相似文献   

7.
Since the seminal work of Caspar and Klug on the structure of the protein containers that encapsulate and hence protect the viral genome, it has been recognised that icosahedral symmetry is crucial for the structural organisation of viruses. In particular, icosahedral symmetry has been invoked in order to predict the surface structures of viral capsids in terms of tessellations or tilings that schematically encode the locations of the protein subunits in the capsids. Whilst this approach is capable of predicting the relative locations of the proteins in the capsids, information on their tertiary structures and the organisation of the viral genome within the capsid are inaccessible. We develop here a mathematical framework based on affine extensions of the icosahedral group that allows us to describe those aspects of the three-dimensional structure of simple viruses. This approach complements Caspar-Klug theory and provides details on virus structure that have not been accessible with previous methods, implying that icosahedral symmetry is more important for virus architecture than previously appreciated.   相似文献   

8.
Influenza A viruses cause a highly contagious respiratory disease in humans and are responsible for periodic widespread epidemics with high mortality rates. The influenza A virus NS1 protein (NS1A) plays a key role in countering host antiviral defense and in virulence. The 73-residue N-terminal domain of NS1A (NS1A-(1-73)) forms a symmetric homodimer with a unique six-helical chain fold. It binds canonical A-form double-stranded RNA (dsRNA). Mutational inactivation of this dsRNA binding activity of NS1A highly attenuates virus replication. Here, we have characterized the unique structural features of the dsRNA binding surface of NS1A-(1-73) using NMR methods and describe the 2.1-A x-ray crystal structure of the corresponding dsRNA binding domain from human influenza B virus NS1B-(15-93). These results identify conserved dsRNA binding surfaces on both NS1A-(1-73) and NS1B-(15-93) that are very different from those indicated in earlier "working models" of the complex between dsRNA and NS1A-(1-73). The combined NMR and crystallographic data reveal highly conserved surface tracks of basic and hydrophilic residues that interact with dsRNA. These tracks are structurally complementary to the polyphosphate backbone conformation of A-form dsRNA and run at an approximately 45 degrees angle relative to the axes of helices alpha2/alpha2'. At the center of this dsRNA binding epitope, and common to NS1 proteins from influenza A and B viruses, is a deep pocket that includes both hydrophilic and hydrophobic amino acids. This pocket provides a target on the surface of the NS1 protein that is potentially suitable for the development of antiviral drugs targeting both influenza A and B viruses.  相似文献   

9.
Penicillium stoloniferum NRRL5267 contains two electrophoretically distinct viruses (PsV-F and PsV-S). An in vivo system was developed to test whether a number of fungal metabolites had antiviral properties on PsV-F replication in O.erties on PsV-F replication in P. stoloniferum. Preliminary results indicated that the mycotoxin patulin can block mycovirus replication. Portions of 48 h mycelium were incubated in the presence of varying levels of patulin, and after an additional 48 h incubation, PsV-F content was measured in E260 units by polyacrylamide gel electrophoresis. Patulin at 11, 16 and 20 mug/mg dry wt mycelia blocked PsV-F replication 26, 61 and 71%, respectively, compared with untreated controls. At these levels, host biomass RNA and protein synthesis were minimally affected. No-proliferating fungal mycelium is capable of continued support of PsV-F replication, which is sensitive to patulin. Apparently, inhibitory doses of patulin stimulated PsV-S replication during this 48 h incubation. The preferential action of patulin may arise from metabolite binding to functional enzymes required for virus replication.  相似文献   

10.
The structure of recombinant virus-like particles of malabaricus grouper nervous necrosis virus (MGNNV), a fish nodavirus isolated from the grouper Epinephelus malabaricus, was determined by electron cryomicroscopy (cryoEM) and three-dimensional reconstruction at 23-A resolution. The cryoEM structure, sequence comparison, and protein fold recognition analysis indicate that the coat protein of MGNNV has two domains resembling those of tomato bushy stunt virus and Norwalk virus, rather than the expected single-domain coat protein of insect nodaviruses. The analysis implies that residues 83 to 216 fold as a beta-sandwich which forms the inner shell of the T=3 capsid and residues 217 to 308 form the trimeric surface protrusions observed in the cryoEM map. The structural similarities between fish nodaviruses and members of the tombusvirus and calicivirus groups provide significant new data for understanding the evolution of the nodavirus family.  相似文献   

11.
Cryoelectron microscopy and three-dimensional computer reconstruction techniques have been used to compare the structures of two types of DNA-free capsids of equine herpesvirus 1 at a resolution of 4.5 nm. "Light" capsids are abortive, whereas "intermediate" capsids are related to maturable intracellular precursors. Their T = 16 icosahedral outer shells, approximately 125 nm in diameter, are indistinguishable and may be described in terms of three layers of density, totalling 15 nm in thickness. The outermost layer consists of protruding portions of both the hexon and the penton capsomers, rising approximately 5 nm above a midlayer of density. The innermost layer, or "floor," is a 4-nm-thick sheet of virtually continuous density except for the orifices of the channels that traverse each capsomer. Hexon protrusions are distinctly hexagonal in shape, and penton protrusions are pentagonal. The structures of the three kinds of hexons (distinguished according to their positions on the surface lattice) are closely similar but differ somewhat in their respective orientations and in the shapes of their channels. The most prominent features of the midlayer are threefold nodules ("triplexes") at the trigonal lattice points. By analogy with other viral capsids, the triplexes may represent trimers of another capsid protein, possibly VP23 (36 kilodaltons [kDa]) or VP26 (12 kDa). Intermediate capsids differ from light capsids, which are empty, in having one or more internal components. In individual images from which the shell structure has been filtered away, these components are seen to have dimensions of 20 to 30 nm but to lack a visible substructure. This material--which is smeared out in the reconstruction, implying that its distribution is not icosahedrally symmetric or necessarily consistent from particle to particle--consists of aggregates of VP22 (46 kDa). From several lines of evidence, we conclude that this protein is located entirely within the capsid shell. These aggregates may be the remnants of morphogenetic cores retained in capsids interrupted in the process of DNA packaging.  相似文献   

12.
Assembly of bacteriophage P22 procapsids has long served as a model for assembly of spherical viruses. Historically, assembly of viruses has been viewed as a non-equilibrium process. Recently alternative models have been developed that treat spherical virus assembly as an equilibrium process. Here we have investigated whether P22 procapsid assembly reactions achieve equilibrium or are irreversibly trapped. To assemble a procapsid-like particle in vitro, pure coat protein monomers are mixed with scaffolding protein. We show that free subunits can exchange with assembled structures, indicating that assembly is a reversible, equilibrium process. When empty procapsid shells (procapsids with the scaffolding protein stripped out) were diluted so that the concentration was below the dissociation constant ( approximately 5 microM) for coat protein monomers, free monomers were detected. The released monomers were assembly-competent; when NaCl was added to metastable partial capsids that were aged for an extended period, the released coat subunits were able to rapidly re-distribute from the partial capsids and form whole procapsids. Lastly, radioactive monomeric coat subunits were able to exchange with the subunits from empty procapsid shells. The data presented illustrate that coat protein monomers are able to dissociate from procapsids in an active state, that assembly of procapsids is consistent with reactions at equilibrium and that the reaction follows the law of mass action.  相似文献   

13.
Structure of an insect virus at 3.0 A resolution   总被引:11,自引:0,他引:11  
We report the first atomic resolution structure of an insect virus determined by single crystal X-ray diffraction. Black beetle virus has a bipartite RNA genome encapsulated in a single particle. The capsid contains 180 protomers arranged on a T = 3 surface lattice. The quaternary organization of the protomers is similar to that observed in the T = 3 plant virus structures. The protomers consist of a basic, crystallographically disordered amino terminus (64 residues), a beta-barrel as seen in other animal and plant virus subunits, an outer protrusion composed predominantly of beta-sheet and formed by three large insertions between strands of the barrel, and a carboxy terminal domain composed of two distorted helices lying inside the shell. The outer surfaces of quasi-threefold related protomers form trigonal pyramidyl protrusions. A cleavage site, located 44 residues from the carboxy terminus, lies within the central cavity of the protein shell. The structural motif observed in BBV (a shell composed of 180 eight-stranded antiparallel beta-barrels) is common to all nonsatellite spherical viruses whose structures have so far been solved. This highly conserved shell architecture suggests a common origin for the coat protein of spherical viruses, while the primitive genome structure of BBV suggests that this insect virus represents an early stage in the evolution of spherical viruses from cellular genes.  相似文献   

14.
15.
16.
Spherical viruses are remarkably well characterized by the Triangulation (T) number developed by Casper and Klug. The T-number specifies how many viral capsid proteins are required to cover the virus, as well as how they are further subdivided into pentamer and hexamer subunits. The T-number however does not constrain the orientations of these proteins within the subunits or dictate where the proteins should place their protruding features. These protrusions often take the form of loops, spires and helices, and are significant because they aid in stability of the capsid as well as recognition by the host organism. Until now there has be no overall understanding of the placement of protrusions for spherical viruses, other than they have icosahedral symmetry. We constructed a set of gauge points based upon the work affine extensions of Keef and Twarock, which have fixed relative angular locations with which to measure the locations of these features. This work adds a new element to our understanding of the geometric arrangement of spherical viral capsid proteins; chiefly that the locations of protruding features are not found stochastically distributed in an icosahedral manner across the viral surface, but instead these features are found only in specific locations along the 15 icosahedral great circles. We have found that this result holds true as the T number and viral capsids size increases, suggesting an underlying geometric constraint on their locations. This is in spite of the fact that the constraints on the pentamers and hexamer orientations change as a function of T-number, as you need to accommodate more hexamers in the same solid angle between pentamers. The existence of this angular constraint of viral capsids suggests that there is a fitness or energetic benefit to the virus placing its protrusions in this manner. This discovery may have profound impacts on identifying and eliminating viral pathogens, understanding evolutionary constraints as well as bioengineering for capsid drug delivery systems. This result also suggests that in addition to biochemical attachment restrictions, there are additional geometric constraints that should be adhered to when modifying protein capsids.  相似文献   

17.
The structure of the T=3 single stranded RNA tymovirus, physalis mottle virus (PhMV), has been determined to 3.8 A resolution. PhMV crystals belong to the rhombohedral space group R 3, with one icosahedral particle in the unit cell leading to 20-fold non-crystallographic redundancy. Polyalanine coordinates of the related turnip yellow mosaic virus (TYMV) with which PhMV coat protein shares 32 % amino acid sequence identity were used for obtaining the initial phases. Extensive phase refinement by real space molecular replacement density averaging resulted in an electron density map that revealed density for most of the side-chains and for the 17 residues ordered in PhMV, but not seen in TYMV, at the N terminus of the A subunits. The core secondary and tertiary structures of the subunits have a topology consistent with the capsid proteins of other T=3 plant viruses. The N-terminal arms of the A subunits, which constitute 12 pentamers at the icosahedral 5-fold axes, have a conformation very different from the conformations observed in B and C subunits that constitute hexameric capsomers with near 6-fold symmetry at the icosahedral 3-fold axes. An analysis of the interfacial contacts between protein subunits indicates that the hexamers are held more strongly than pentamers and hexamer-hexamer contacts are more extensive than pentamer-hexamer contacts. These observations suggest a plausible mechanism for the formation of empty capsids, which might be initiated by a change in the conformation of the N-terminal arm of the A subunits. The structure also provides insights into immunological and mutagenesis results. Comparison of PhMV with the sobemovirus, sesbania mosaic virus reveals striking similarities in the overall tertiary fold of the coat protein although the capsid morphologies of these two viruses are very different.  相似文献   

18.
19.
20.
Assembly intermediates of icosahedral viruses are usually transient and are difficult to identify. In the present investigation, site-specific and deletion mutants of the coat protein gene of physalis mottle tymovirus (PhMV) were used to delineate the role of specific amino acid residues in the assembly of the virus and to identify intermediates in this process. N-terminal 30, 34, 35 and 39 amino acid deletion and single C-terminal (N188) deletion mutant proteins of PhMV were expressed in Escherichia coli. Site-specific mutants H69A, C75A, W96A, D144N, D144N-T151A, K143E and N188A were also constructed and expressed. The mutant protein lacking 30 amino acid residues from the N terminus self-assembled to T=3 particles in vivo while deletions of 34, 35 and 39 amino acid residues resulted in the mutant proteins that were insoluble. Interestingly, the coat protein (pR PhCP) expressed using pRSET B vector with an additional 41 amino acid residues at the N terminus also assembled into T=3 particles that were more compact and had a smaller diameter. These results demonstrate that the amino-terminal segment is flexible and either the deletion or addition of amino acid residues at the N terminus does not affect T=3 capsid assembly. In contrast, the deletion of even a single residue from the C terminus (PhN188Delta1) resulted in capsids that were unstable. These capsids disassembled to a discrete intermediate with a sedimentation coefficent of 19.4 S. However, the replacement of C-terminal asparagine 188 by alanine led to the formation of stable capsids. The C75A and D144N mutant proteins also assembled into capsids that were as stable as the pR PhCP, suggesting that C75 and D144 are not crucial for the T=3 capsid assembly. pR PhW96A and pR PhD144N-T151A mutant proteins failed to form capsids and were present as heterogeneous aggregates. Interestingly, the pR PhK143E mutant protein behaved in a manner similar to the C-terminal deletion protein in forming unstable capsids. The intermediate with an s value of 19.4 S was the major assembly product of pR PhH69A mutant protein and could correspond to a 30mer. It is possible that the assembly or disassembly is arrested at a similar stage in pR PhN188Delta1, pR PhH69A and pR PhK143E mutant proteins.  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