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.
Chilo iridescent virus (CIV) is a large (∼ 1850 Å diameter) insect virus with an icosahedral, T = 147 capsid, a double-stranded DNA (dsDNA) genome, and an internal lipid membrane. The structure of CIV was determined to 13 Å resolution by means of cryoelectron microscopy (cryoEM) and three-dimensional image reconstruction. A homology model of P50, the CIV major capsid protein (MCP), was built based on its amino acid sequence and the structure of the homologous Paramecium bursaria chlorella virus 1 Vp54 MCP. This model was fitted into the cryoEM density for each of the 25 trimeric CIV capsomers per icosahedral asymmetric unit. A difference map, in which the fitted CIV MCP capsomers were subtracted from the CIV cryoEM reconstruction, showed that there are at least three different types of minor capsid proteins associated with the capsomers outside the lipid membrane. “Finger” proteins are situated at many, but not all, of the spaces between three adjacent capsomers within each trisymmetron, and “zip” proteins are situated between sets of three adjacent capsomers at the boundary between neighboring trisymmetrons and pentasymmetrons. Based on the results of segmentation and density correlations, there are at least eight finger proteins and three dimeric and two monomeric zip proteins in one asymmetric unit of the CIV capsid. These minor proteins appear to stabilize the virus by acting as intercapsomer cross-links. One transmembrane “anchor” protein per icosahedral asymmetric unit, which extends from beneath one of the capsomers in the pentasymmetron to the internal leaflet of the lipid membrane, may provide additional stabilization for the capsid. These results are consistent with the observations for other large, icosahedral dsDNA viruses that also utilize minor capsid proteins for stabilization and for determining their assembly.  相似文献   

2.
Bacteriophage G4 and φX174 are members of the Microviridae family. The degree of similarity of the structural proteins ranges from 66% identity of the F protein to 40% identity of the G protein. The atomic structure of the φX174 virion had previously been determined by X-ray crystallography. Bacteriophage G4 procapsids, consisting of the structural proteins F, G, D, B, H, and small traces of J but no DNA, were set up for crystallization. However, the resultant crystals were of degraded procapsid particles, which had lost the assembly scaffolding proteins D and B, resulting in particles that resembled empty virions.The structure of the degraded G4 procapsid has been determined to 3.0 Å resolution. The particles crystallized in the hexagonal space groupP6322 with unit cell dimensionsa=b=414.2(5) Å andc=263.0(3) Å. The diffraction data were collected at the Cornell High Energy Synchrotron Source (CHESS) on film and image plates using oscillation photography. Packing considerations indicated there were two particles per unit cell. A self-rotation function confirmed that the particles were positioned on 32 point group special positions in the unit cell. Initial phases were calculated to 6 Å resolution, based on the known φX174 virion model. Phase information was then extended in steps to 3.0 Å resolution by molecular replacement electron density modification and particle envelope generation.The resulting electron density map was readily interpretable in terms of the F and G polypeptides, as occur in the mature capsid of φX174. In a few regions of the electron density map there were inconsistencies between the density and the published amino acid sequence. Redetermining the amino acid sequence confirmed that the density was correct. The r.m.s. deviation between the Cαbackbone of the mature capsid of φX174 and the degraded G4 procapsid was 0.36 Å for the F protein and 1.38 Å for the G protein. This is consistent with the greater conservation of the F protein compared to the G protein sequences among members of the Microviridae family. Functionally important features between φX174 and G4 had greater conservation.Calcium ions (Ca2 +) were shown to bind to G4 at a general site located near the icosahedral 3-fold axis on the F protein capsid, equivalent to sites found previously in φX174. Binding of Ca2 +also caused the ordering of the conserved region of the DNA binding protein J, which was present in the degraded procapsid particle in the absence of DNA.  相似文献   

3.
Single-stranded RNA viruses package their genomes into capsids enclosing fixed volumes. We assayed the ability of bacteriophage MS2 coat protein to package large, defined fragments of its genomic, single-stranded RNA. We show that the efficiency of packaging into a T = 3 capsid in vitro is inversely proportional to RNA length, implying that there is a free-energy barrier to be overcome during assembly. All the RNAs examined have greater solution persistence lengths than the internal diameter of the capsid into which they become packaged, suggesting that protein-mediated RNA compaction must occur during assembly. Binding ethidium bromide to one of these RNA fragments, which would be expected to reduce its flexibility, severely inhibited packaging, consistent with this idea. Cryo-EM structures of the capsids assembled in these experiments with the sub-genomic RNAs show a layer of RNA density beneath the coat protein shell but lack density for the inner RNA shell seen in the wild-type virion. The inner layer is restored when full-length virion RNA is used in the assembly reaction, implying that it becomes ordered only when the capsid is filled, presumably because of the effects of steric and/or electrostatic repulsions. The cryo-EM results explain the length dependence of packaging. In addition, they show that for the sub-genomic fragments the strongest ordered RNA density occurs below the coat protein dimers forming the icosahedral 5-fold axes of the capsid. There is little such density beneath the proteins at the 2-fold axes, consistent with our model in which coat protein dimers binding to RNA stem-loops located at sites throughout the genome leads to switching of their preferred conformations, thus regulating the placement of the quasi-conformers needed to build the T = 3 capsid. The data are consistent with mutual chaperoning of both RNA and coat protein conformations, partially explaining the ability of such viruses to assemble so rapidly and accurately.  相似文献   

4.
5.
A trans-encapsidation assay was established to study the specificity of picornavirus RNA encapsidation. A poliovirus replicon with the luciferase gene replacing the capsid protein-coding region was coexpressed in transfected HeLa cells with capsid proteins from homologous or heterologous virus. Successful trans-encapsidation resulted in assembly and production of virions whose replication, upon subsequent infection of HeLa cells, was accompanied by expression of luciferase activity. The amount of luciferase activity was proportional to the amount of trans-encapsidated virus produced from the cotransfection. When poliovirus capsid proteins were supplied in trans, >2 × 106 infectious particles/ml were produced. When coxsackievirus B3, human rhinovirus 14, mengovirus, or hepatitis A virus (HAV) capsid proteins were supplied in trans, all but HAV showed some encapsidation of the replicon. The overall encapsidation efficiency of the replicon RNA by heterologous capsid proteins was significantly lower than when poliovirus capsid was used. trans-encapsidated particles could be completely neutralized with specific antisera against each of the donor virus capsids. The results indicate that encapsidation is regulated by specific viral nucleic acid and protein sequences.  相似文献   

6.
7.
Although the plasma membrane is the terminal destination for glycosylphosphatidylinositol (GPI) proteins in higher eukaryotes, cell wall-attached GPI proteins (GPI-CWPs) are found in many fungal species. In yeast, some of the cis-requirements directing localization of GPI proteins to the plasma membrane or cell wall are now understood. However, it remains to be determined how Aspergillus fumigatus, an opportunistic fungal pathogen, signals, and sorts GPI proteins to either the plasma membrane or the cell wall. In this study, chimeric green fluorescent proteins (GFPs) were constructed as fusions with putative C-terminal GPI signal sequences from A. fumigatus Mp1p, Gel1p, and Ecm33p, as well as site-directed mutations thereof. By analyzing cellular localization of chimeric GFPs using Western blotting, electron microscopy, and fluorescence microscopy, we showed that, in contrast to yeast, a single Lys residue at the ω-1 or ω-2 site alone could retain GPI-anchored GFP in the plasma membrane. Although the signal for cell wall distribution has not been identified yet, it appeared that the threonine/serine-rich region at the C-terminal half of AfMp1 was not required for cell wall distribution. Based on our results, the cis-requirements directing localization of GPI proteins in A. fumigatus are different from those in yeast.  相似文献   

8.
The orientation and location of the 240 hexons comprising the outer protein shell of adenovirus have been determined. Electron micrographs of the capsid and its fragments were inspected for the features of hexon known from the X-ray crystallographic model as described in the accompanying paper. A capsid model is proposed with each facet comprising a small p3 net of 12 hexons, arranged as a triangular sextet with three outer hexon pairs. The sextet is centrally placed about the icosahedral threefold axis, with its edges parallel to those of the facet. The outer pairs project over the facet edges on one side of the icosahedral twofold axes at each edge. The model capsid is defined by the underlying icosahedron, of edge 445 A, upon which hexons are arranged. The hexons are thus bounded by icosahedra with insphere radii of 336 A and 452 A. A quartet of hexons forms the asymmetric unit of an icosahedral hexon shell, which can be closed by the addition of pentons at the 12 vertices. Considering the hexon trimer as a complex structure unit, its interactions in the four topologically distinct environments are very similar, with conservation of at least two-thirds of the inter-hexon bonding. The crystal-like construction explains the flat facets and sharp edges characteristic of adenovirus. Larger "adenovirus-like" capsids of any size could be formed using only one additional topologically different environment. The construction of adenovirus illustrates how an impenetrable protein shell can be formed, with highly conserved intermolecular bonding, by using the geometry of an oligomeric structure unit and symmetry additional to that of the icosahedral point group. This contrasts with the manner suggested by Caspar & Klug (1962), in which the polypeptide is the structure unit, and for which the number of possible bonding configurations required of a structure unit tends to infinity as the continuously curved capsid increases in size. The known structures of polyoma and the plant viruses with triangulation number equal to 3 are evaluated in terms of hexamer-pentamer packing, and evidence is presented for the existence of larger subunits than the polypeptide in both cases. It is suggested that spontaneous assembly can occur only when exact icosahedral symmetry relates structure units or sub-assemblies, which would themselves have been formed by self-limiting closed interactions. Without such symmetry, the presence of scaffolding proteins or nucleic acid is necessary to limit aggregation.  相似文献   

9.
BackgroundThe structures of proteins are intimately related to their functions. Significant efforts have been dedicated to the structural investigation of proteins, mainly those of purified proteins in in vitro environments. Proteins function in living cells and thus protein structures must be regulated by interactions with various molecules, some of which participate in reaction networks, depending on the states, conditions, or actions of the cell. Therefore, it is very important to understand the structural behavior of proteins in living cells.MethodsSingle-molecule Förster resonance energy transfer (smFRET) measurements were conducted using the alternative laser excitation (ALEX) technique. smFRET distributions of cytosolic Rapidly Accelerated Fibrosarcoma (RAF) proteins in living HeLa cells were obtained with exclusion of the negative effects of photobleached fluorophores and incompletely labeled proteins on smFRET.ResultssmFRET histograms of wildtype (wt) RAF in live cells exhibited two major peaks, whereas that of the S621A mutant, which has been thought to have an expanded structure, was almost single-peaked. A population shift involving the peaks for wt RAF was detected upon epidermal growth factor stimulation. Spontaneous transitions between the conformational states corresponding to the two peaks were also detected using the FRET-two-channel kernel-based density distribution estimator method in comparison to static double-stranded DNA samples.ConclusionsCytosolic CRAF has at least three conformational states; in addition to the closed and open forms, the fully-open form was distinctly specified. Based on the results, we propose a speculative structural model for CRAF.General significanceStructural distribution and changes to proteins in live cells as a result of intracellular interactions were successfully identified. smFRET using ALEX is applicable to any other cytosolic proteins.  相似文献   

10.
The Staphylococcus aureus pathogenicity island SaPI1 carries the gene for the toxic shock syndrome toxin (TSST-1) and can be mobilized by infection with S. aureus helper phage 80α. SaPI1 depends on the helper phage for excision, replication and genome packaging. The SaPI1-transducing particles comprise proteins encoded by the helper phage, but have a smaller capsid commensurate with the smaller size of the SaPI1 genome. Previous studies identified only 80α-encoded proteins in mature SaPI1 virions, implying that the presumptive SaPI1 capsid size determination function(s) must act transiently during capsid assembly or maturation. In this study, 80α and SaPI1 procapsids were produced by induction of phage mutants lacking functional 80α or SaPI1 small terminase subunits. By cryo-electron microscopy, these procapsids were found to have a round shape and an internal scaffolding core. Mass spectrometry was used to identify all 80α-encoded structural proteins in 80α and SaPI1 procapsids, including several that had not previously been found in the mature capsids. In addition, SaPI1 procapsids contained at least one SaPI1-encoded protein that has been implicated genetically in capsid size determination. Mass spectrometry on full-length phage proteins showed that the major capsid protein and the scaffolding protein are N-terminally processed in both 80α and SaPI1 procapsids.  相似文献   

11.
A model of the RNA of tobacco mosaic virus has been built using computer model-building techniques. The model has good stereochemistry, and fits the electron density map of the virus obtained by fiber diffraction methods considerably better than did earlier models. The three sugar rings in the asymmetric unit all have the A (3′-endo) conformation, One of the bases is in the syn conformation, a conformation observed only rarely in nucleic acid structures.  相似文献   

12.
While many aspects of plant cell wall polymer structure are known, their spatial and temporal distribution within the stem are not well understood. Here, we studied vascular system and fiber development, which has implication for both biofuel feedstock conversion efficiency and crop yield. The subject of this study, Brachypodium distachyon, has emerged as a grass model for food and energy crop research. Here, we conducted our investigation using B. distachyon by applying various histological approaches and Fourier transform infrared spectroscopy to the stem internode from three key developmental stages. While vascular bundle size and number did not change over time, the size of the interfascicular region increased dramatically, as did cell wall thickness. We also describe internal stem internode anatomy and demonstrate that lignin deposition continues after crystalline cellulose and xylan accumulation ceases. The vascular bundle anatomy of B. distachyon appears to be highly similar to domesticated grasses. While the arrangement of bundles within the stem is highly variable across grasses, B. distachyon appears to be a suitable model for the rind of large C4 grass crops. A better understanding of growth and various anatomical and cell wall features of B. distachyon will further our understanding of plant biomass accumulation processes.  相似文献   

13.
Small-angle neutron scattering experiments have been performed on the tubular bottom component of Alfalfa mosaic virus (AMV) and the “30 S” particle (a quasispherical reassembled AMV coat protein particle) with the aim of determining the internal structure of the virus. Scattering curves were obtained out to a resolution of 150A??1 at a number of H2O/2H2O ratios and were analysed using a model fitting technique. This involves calculating the scattering intensity due to a parameterised distribution of scattering density representing the particle and comparing this to the experimental data after taking into account the effect of instrumental smearing. The use of the contrast variation method enables the internal consistency of the model to be well tested.Three models are used in an attempt to explain the scattering curve of the 30 S particle. A single homogeneous shell is shown to be inadequate and two other models introducing the presumed T = 1 icosahedral symmetry of the particle are presented and discussed. The most satisfactory of these consists of 60 spherical monomers of radius 19 Å symmetrically placed in pairs about the 2-fold icosahedral positions.The analysis of the bottom component data has yielded a low resolution model for the virus, which is shown to be consistent with its composition as given by earlier physico-chemical measurements. In the model the RNA is uniformly packed throughout the interior of the capsid (which is cylindrical with hemispherical ends) out to a radius of about 65 Å and with a packing fraction of 20%. Within the limitations of an homogeneous shell model, the protein capsid has an outer radius of 94 Å and thickness of 23 Å, but arguments are presented based on the marked lattice structure of the cylindrical capsid and the analysis of the scattering data of the 30 S particle, that this model underestimates the thickness of the protein shell and that it in fact makes contact with the RNA at about 65 Å.  相似文献   

14.
15.
The structure of the Leviviridae bacteriophage φCb5 virus-like particle has been determined at 2.9 Å resolution and the structure of the native bacteriophage φCb5 at 3.6 Å. The structures of the coat protein shell appear to be identical, while differences are found in the organization of the density corresponding to the RNA. The capsid is built of coat protein dimers and in shape corresponds to a truncated icosahedron with T = 3 quasi-symmetry. The capsid is stabilized by four calcium ions per icosahedral asymmetric unit. One is located at the symmetry axis relating the quasi-3-fold related subunits and is part of an elaborate network of hydrogen bonds stabilizing the interface. The remaining calcium ions stabilize the contacts within the coat protein dimer. The stability of the φCb5 particles decreases when calcium ions are chelated with EDTA. In contrast to other leviviruses, φCb5 particles are destabilized in solution with elevated salt concentration. The model of the φCb5 capsid provides an explanation of the salt-induced destabilization of φCb5, since hydrogen bonds, salt bridges and calcium ions have important roles in the intersubunit interactions.Electron density of three putative RNA nucleotides per icosahedral asymmetric unit has been observed in the φCb5 structure. The nucleotides mediate contacts between the two subunits forming a dimer and a third subunit in another dimer. We suggest a model for φCb5 capsid assembly in which addition of coat protein dimers to the forming capsid is facilitated by interaction with the RNA genome. The φCb5 structure is the first example in the levivirus family that provides insight into the mechanism by which the genome-coat protein interaction may accelerate the capsid assembly and increase capsid stability.  相似文献   

16.
Diffraction of X-rays is recorded from barium stearate multilayer systems with from 2 to 60 double layers or unit cells. The generalized Patterson function P′(x) is calculated by an integral Fourier transform of observed intensity data from a specimen containing only two unit cells. The Patterson function P0(x) of a single unit cell is determined from P′(x) and the electron density distribution of a bimolecular leaflet is obtained by a deconvolution procedure of P0(x) after Hosemann and Bagchi. The electron density distribution is also calculated independently by a conventional Fourier synthesis with an experimentally established set of phases. The results of the two methods are consistent and fit a physical model of the bimolecular leaflet. A direct analysis, therefore, can be performed if diffraction is observed from multilayer systems with a small number of unit cells.  相似文献   

17.
Physical and chemical parameters of iridescent virus type 29, isolated from the mealworm, Tenebrio molitor, have been analyzed. The icosahedral capsid is 130–135 nm in diameter and is surrounded by a fringe of coarse filaments. The virus has a buoyant density in CsCl of 1.31 g cm?3 and contains 20 to 25 structural proteins as analyzed by isoelectric focusing and SDS-polyacrylamide gel electrophoresis. The DNA has a buoyant density in CsCl of 1.6874 g cm?3 indicating a G + C content of approximately 28%. The lipid components of this virus differ from those of the host cell; the virus contains about 80% cardiolipin and 20% phosphatidyl choline.  相似文献   

18.
Coxsackievirus A7 (CAV7) is a rarely detected and poorly characterized serotype of the Enterovirus species Human enterovirus A (HEV-A) within the Picornaviridae family. The CAV7-USSR strain has caused polio-like epidemics and was originally thought to represent the fourth poliovirus type, but later evidence linked this strain to the CAV7-Parker prototype. Another isolate, CAV7-275/58, was also serologically similar to Parker but was noninfectious in a mouse model. Sequencing of the genomic region encoding the capsid proteins of the USSR and 275/58 strains and subsequent comparison with the corresponding amino acid sequences of the Parker strain revealed that the Parker and USSR strains are nearly identical, while the 275/58 strain is more distant. Using electron cryomicroscopy and three-dimensional image reconstruction, the structures of the CAV7-USSR virion and empty capsid were resolved to 8.2-Å and 6.1-Å resolutions, respectively. This is one of the first detailed structural analyses of the HEV-A species. Using homology modeling, reconstruction segmentation, and flexible fitting, we constructed a pseudoatomic T = 1 (pseudo T = 3) model incorporating the three major capsid proteins (VP1 to VP3), addressed the conformational changes of the capsid and its constituent viral proteins occurring during RNA release, and mapped the capsid proteins'' variable regions to the structure. During uncoating, VP4 and RNA are released analogously to poliovirus 1, the interfaces of VP2 and VP3 are rearranged, and VP1 rotates. Variable regions in the capsid proteins were predicted to map mainly to the surface of VP1 and are thus likely to affect the tropism and pathogenicity of CAV7.  相似文献   

19.
Bacteriophages are involved in many aspects of the spread and establishment of virulence factors in Staphylococcus aureus, including the mobilization of genetic elements known as S. aureus pathogenicity islands (SaPIs), which carry genes for superantigen toxins and other virulence factors. SaPIs are packaged into phage-like transducing particles using proteins supplied by the helper phage. We have used cryo-electron microscopy and icosahedral reconstruction to determine the structures of the procapsid and the mature capsid of 80α, a bacteriophage that can mobilize several different SaPIs. The 80α capsid has T = 7 icosahedral symmetry with the capsid protein organized into pentameric and hexameric clusters that interact via prominent trimeric densities. The 80α capsid protein was modeled based on the capsid protein fold of bacteriophage HK97 and fitted into the 80α reconstructions. The models show that the trivalent interactions are mediated primarily by a 22-residue β hairpin structure called the P loop that is not found in HK97. Capsid expansion is associated with a conformational switch in the spine helix that is propagated throughout the subunit, unlike the domain rotation mechanism in phage HK97 or P22.  相似文献   

20.

Background

The mouse Fv1 (friend virus) susceptibility gene inhibits the development of the murine leukaemia virus (MLV) by interacting with its capsid (CA) protein. As no structures are available for these proteins we have constructed molecular models based on distant sequence similarity to other retroviral capsid proteins.

Results

Molecular models were constructed for the amino terminal domains of the probable capsid-like structure for the mouse Fv1 gene product and the capsid protein of the MLV. The models were based on sequence alignments with a variety of other retrovirus capsid proteins. As the sequence similarity of these proteins with MLV and especially Fv1 is very distant, a threading method was employed that incorporates predicted secondary structure and multiple sequence information. The resulting models were compared with equivalent models constructed using the sequences of the capsid proteins of known structure.

Conclusions

These comparisons suggested that the MLV model should be accurate in the core but with significant uncertainty in the loop regions. The Fv1 model may have some additional errors in the core packing of its helices but the resulting model gave some support to the hypothesis that it adopts a capsid-like structure.
  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