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.
The three-dimensional structure of capsid of Aedes albopictus C6/36 densovirus was determined to 14-(A) resolution by electron cryomicroscopy and computer reconstruction. The triangulation number of the capsid is 1. There are 12 holes in each triangular face and a spike on each 5-fold vertex. The validity of the capsid and nucleic acid densities in the reconstructions was discussed.  相似文献   

2.
The three-dimensional structure of capsid ofAedes albopictus C6/36 densovirus was determined to 14-Å resolution by electron cryomicroscopy and computer reconstruction. The triangulation number of the capsid is 1. There are 12 holes in each triangular face and a spike on each 5-fold vertex. The validity of the capsid and nucleic acid densities in the reconstructions was discussed.  相似文献   

3.
The three-dimensional structure of Nudaurelia capensis beta virus (N beta V) was reconstructed to 3.2-nm resolution from images of frozen-hydrated virions. The distinctly icosahedral capsid (approximately 40-nm diameter) contains 240 copies of a single 61-kDa protein subunit arranged with T = 4 lattice symmetry. The outer surface of unstained virions compares remarkably well with that previously observed in negatively stained specimens. Inspection of the density map, volume estimates, and model building experiments indicate that each subunit consists of two distinct domains. The large domain (approximately 40 kDa) has a cylindrical shape, approximately 4-nm diameter by approximately 4-nm high, and associates with two large domains of neighboring subunits to form a Y-shaped trimeric aggregate in the outer capsid surface. Four trimers make up each of the 20 planar faces of the capsid. Small domains (approximately 21 kDa) presumably associate at lower radii (approximately 13-16.5 nm) to form a contiguous, non-spherical shell. A T = 4 model, constructed from 80 trimers of the common beta-barrel core motif (approximately 20 kDa) found in many of the smaller T = 3 and pseudo T = 3 viruses, fits the dimensions and features seen in the N beta V reconstruction, suggesting that the contiguous shell of N beta V may be formed by intersubunit contacts between small domains having that motif. The small (approximately 1800 kDa), ssRNA genome is loosely packed inside the capsid with a low average density.  相似文献   

4.
The minute virus of mice (MVM) provides a simple model for the dissection of the molecular determinants of the self-assembly, stability, and dynamics of a biological supramolecular complex. MVM assembly involves the trimerization of capsid subunits in the cytoplasm; trimers are transported to the nucleus, where they suffer a conformational change and are made competent for capsid formation. Our previous study revealed that capsid assembly from trimers is dependent on stronger intertrimer interactions that are equally spaced in an equatorial belt surrounding each trimer. We have now targeted the interfaces between monomers within each trimer to identify the molecular determinants of trimerization and the rearrangement needed for capsid assembly. Twenty-eight amino acid residues per monomer were individually mutated to alanine to remove most of the stronger intersubunit interactions. The effects on trimer and capsid assembly and virus infectivity in cells were analyzed. No side chain was individually required for trimer assembly in the cytoplasm; in contrast, half of them were required to make the trimers competent for nuclear capsid assembly, even though none was close to intertrimer interfaces. These critical side chains are conserved and participate in extensive hydrophobic contacts, buried hydrogen bonds, or salt bridges between subunits. This study on MVM capsid assembly reveals that: (i) trimerization is a robust process, insensitive to removal of individual intersubunit interactions; and (ii) the rearrangement of the trimer intermediate required for capsid assembly is a global process that depends on the establishment of many interactions along the protein-protein interfaces within each trimer.  相似文献   

5.
HK97 is an exceptionally amenable system for characterizing major conformational changes associated with capsid maturation in double-stranded DNA bacteriophage. HK97 undergoes a capsid expansion of ∼ 20%, accompanied by major subunit rearrangements during genome packaging. A previous 3.44-Å-resolution crystal structure of the mature capsid Head II and cryo-electron microscopy studies of other intermediate expansion forms of HK97 suggested that, primarily, rigid-body movements facilitated the maturation process. We recently reported a 3.65-Å-resolution structure of the preexpanded particle form Prohead II (P-II) and found that the capsid subunits undergo significant refolding and twisting of the tertiary structure to accommodate expansion. The P-II study focused on major twisting motions in the P-domain and on refolding of the spine helix during the transition. Here we extend the crystallographic comparison between P-II and Head II, characterizing the refolding events occurring in each of the four major domains of the capsid subunit and their effect on quaternary structure stabilization. In addition, hydrogen/deuterium exchange, coupled to mass spectrometry, was used to characterize the structural dynamics of three distinct capsid intermediates: P-II, Expansion Intermediate, and the nearly mature Head I. Differences in the solvent accessibilities of the seven quasi-equivalent capsid subunits, attributed to differences in secondary and quaternary structures, were observed in P-II. Nearly all differences in solvent accessibility among subunits disappear after the first transition to Expansion Intermediate. We show that most of the refolding is coupled to this transformation, an event associated with the transition from asymmetric to symmetric hexamers.  相似文献   

6.
BACKGROUND: The dsDNA bacteriophage PRD1 has a membrane inside its icosahedral capsid. While its large size (66 MDa) hinders the study of the complete virion at atomic resolution, a 1.65-A crystallographic structure of its major coat protein, P3, is available. Cryo-electron microscopy (cryo-EM) and three-dimensional reconstruction have shown the capsid at 20-28 A resolution. Striking architectural similarities between PRD1 and the mammalian adenovirus indicate a common ancestor. RESULTS: The P3 atomic structure has been fitted into improved cryo-EM reconstructions for three types of PRD1 particles: the wild-type virion, a packaging mutant without DNA, and a P3-shell lacking the membrane and the vertices. Establishing the absolute EM scale was crucial for an accurate match. The resulting "quasi-atomic" models of the capsid define the residues involved in the major P3 interactions, within the quasi-equivalent interfaces and with the membrane, and show how these are altered upon DNA packaging. CONCLUSIONS: The new cryo-EM reconstructions reveal the structure of the PRD1 vertex and the concentric packing of DNA. The capsid is essentially unchanged upon DNA packaging, with alterations limited to those P3 residues involved in membrane contacts. These are restricted to a few of the N termini along the icosahedral edges in the empty particle; DNA packaging leads to a 4-fold increase in the number of contacts, including almost all copies of the N terminus and the loop between the two beta barrels. Analysis of the P3 residues in each quasi-equivalent interface suggests two sites for minor proteins in the capsid edges, analogous to those in adenovirus.  相似文献   

7.
Gene Order of the Poliovirus Capsid Proteins   总被引:17,自引:14,他引:3       下载免费PDF全文
Two methods were used to determine the genetic map of the poliovirus capsid proteins. The first method uses pactamycin, a drug which selectively inhibits the initiation of protein synthesis and causes a change in the relative amounts of capsid proteins synthesized. This differential effect on each of the capsid proteins is interpreted as indicating the relative distance of each protein from the initiation site of protein synthesis. The second method involves an analysis of coat precursor molecules released from polyribosomes after a series of short pulses of different length terminated by addition of emetine, a drug which stops all protein synthesis almost immediately after its addition. As the pulse length is increased, each of the capsid proteins within the precursor gains radioactivity with different kinetics. From these kinetics, it is possible to determine the gene order of the capsid proteins within the precursor as well as a rate of protein synthesis. Both methods indicate a gene order for the region of the ribonucleic acid coding for the capsid proteins as (5' --> 3') VP 4 - VP 2 - VP 3 - VP 1.  相似文献   

8.
Ceres P  Zlotnick A 《Biochemistry》2002,41(39):11525-11531
Hepatitis B virus (HBV) is an enveloped DNA virus with a spherical capsid (or core). The capsid is constructed from 120 copies of the homodimeric capsid protein arranged with T = 4 icosahedral symmetry. We examined in vitro assembly of purified E. coli expressed HBV capsid protein. After equilibration, concentrations of capsid and dimer were evaluated by size exclusion chromatography. The extent of assembly increased as temperature and ionic strength increased. The concentration dependence of capsid assembly conformed to the equilibrium expression: K(capsid) = [capsid]/[dimer](120). Given the known geometry for HBV capsids and dimers, the per capsid assembly energy was partitioned into energy per subunit-subunit contact. We were able to make three major conclusions. (i) Weak interactions (from -2.9 kcal/mol at 21 degrees C in low salt to -4.4 kcal/mol at 37 degrees C in high salt) at each intersubunit contact result in a globally stable capsid; weak intersubunit interactions may be the basis for the phenomenon of capsid breathing. (ii) HBV assembly is characterized by positive enthalpy and entropy. The reaction is entropy-driven, consistent with the largely hydrophobic contacts found in the crystal structure. (iii) Increasing NaCl concentration increases the magnitude of free energy, enthalpy, and entropy, as if ionic strength were increasing the amount of hydrophobic surface buried by assembly. This last point leads us to suggest that salt acts by inducing a conformational change in the dimer from an assembly-inactive form to an assembly-active form. This model of conformational change linked to assembly is consistent with immunological differences between dimer and capsid.  相似文献   

9.
Hepatitis B virus morphogenesis is accompanied by the production and release of non‐enveloped capsids/nucleocapsids. Capsid particles are formed inside the cell cytosol by multimerization of core protein subunits and ultimately exported in an uncommon coatless state. Here, we investigated potential roles of Rab GTPases in capsid formation and trafficking by using RNA interference and overexpression studies. Naked capsid release does not require functions of the endosome‐associated Rab5, Rab7 and Rab27 proteins, but depends on functional Rab33B, a GTPase participating in autophagosome formation via interaction with the Atg5‐Atg12/Atg16L1 complex. During capsid formation, Rab33B acts in conjunction with its effector, as silencing of Atg5, Atg12 and Atg16L1 also impaired capsid egress. Analysis of capsid maturation steps revealed that Rab33B and Atg5/12/16L1 are required for proper particle assembly and/or stability. In support, the capsid protein was found to interact with Atg5 and colocalize with Atg5/12/16L1, implicating that autophagy pathway functions are involved in capsid biogenesis. However, a complete and functional autophagy pathway is dispensable for capsid release, as judged by knockdown analysis of Atg8/LC3 family members and pharmaceutical ablation of canonical autophagy. Experiments aimed at analysing the capsid release‐stimulating activity of the Alix protein provide further evidence for a link between capsid formation and autophagy.  相似文献   

10.
UL25 is one of seven herpes simplex virus-encoded proteins involved specifically in DNA encapsidation. Its role appears to be to stabilize the capsid so that DNA is prevented from escaping once it has entered. To clarify the function of UL25, we have examined capsids with the goal of defining where it is located. Analysis of trypsin-treated capsids showed that UL25 is sensitive to cleavage like other proteins such as the major capsid and portal proteins that are exposed on the capsid surface. Internal proteins such as the scaffolding protein and protease were not affected under the same experimental conditions. Capsids were also examined by electron microscopy after staining with gold-labeled antibody specific for UL25. Images of stained capsids demonstrated that most labeled sites (71% in C capsids) were at capsid vertices, and most stained C capsids had label at more than one vertex. A quantitative immunoblotting method showed that the capsid contents of UL25 were 56, 20, and 75 copies per capsid in A, B, and C capsids, respectively. Finally, soluble UL25 protein was found to bind in vitro to purified capsids lacking it. The amount of bound UL25 corresponded to the amount present in B capsids, and bound UL25 was found by immunoelectron microscopy to be located predominantly at the capsid vertices. The results are interpreted to suggest that five UL25 molecules are found at or near each of the capsid vertices, where they are exposed on the capsid surface. Exposure on the surface is consistent with the view that UL25 is added to the capsid as DNA is packaged or during late stages of the packaging process.  相似文献   

11.
Cell-free translation of the Semliki Forest virus-specific 26S RNA yielded primarily capsid protein. After treatment of the protein synthesising reaction with 25 mM EDTA, the capsid protein cosedimented with the large ribosomal subunit in sucrose gradients, and banded with the subunit at a density of 1.54 gm/cm3 in CsCl. Exposure to 0.5 M KCl released the protein from the subunit. Similar binding of the virus capsid protein to the large ribosomal subunit has been observed in infected HeLa cells, although its function is not clear. The nonstructural proteins, which are the major products translated from the virion 42S RNA, did not associate with sedimenting structures.  相似文献   

12.
I Singh  A Helenius 《Journal of virology》1992,66(12):7049-7058
The mechanism by which Semliki Forest virus nucleocapsids are uncoated was analyzed in living cells and in vitro. In BHK-21 cells, uncoating occurred with virtually complete efficiency within 1 to 2 min after the nucleocapsids entered the cytoplasm. It was inhibited by monensin, which blocks nucleocapsid penetration from endosomes. As previously shown for Sindbis virus (G. Wengler and G. Wengler, Virology 134:435-442, 1984), the capsid proteins from incoming nucleocapsids became associated with ribosomes. The ribosome-bound capsid proteins were distributed throughout the cytoplasm, while the viral RNA remained associated with vacuolar membranes. Using purified nucleocapsids and ribosomes in vitro, we established that ribosomes alone were sufficient for uncoating. Their role was to release the capsid proteins from nucleocapsids and irreversibly sequester them, in a process independent of energy and translation. The process was stoichiometric rather than catalytic, with a maximum of three to six capsid proteins bound to each ribosome. More than 80% of the capsid proteins could thus be removed from the viral RNA, resulting in the formation of nucleocapsid remnants whose sedimentation coefficients progressively decreased from 140S to 80S as uncoating proceeded.  相似文献   

13.
The structures of the hexavalent capsomers of herpes simplex virus type 2 were analyzed by negative staining electron microscopy of capsomer patches derived from partially disrupted nucleocapsids. Optimally computer-averaged images were formed for each of the three classes of capsomer distinguished by their respective positions on the surface of the icosahedral capsid with a triangulation number of 16; in projection, each capsomer exhibited unequivocal sixfold symmetry. According to correspondence analysis of our set of capsomer images, no significant structural differences were detected among the three classes of capsomers, as visualized under these conditions. Taking into account information from images of freeze-dried, platinum-shadowed nucleocapsid fragments, it was established that each hexavalent capsomer is a hexamer of the 155-kilodalton major capsid protein. The capsomer has the form of a sixfold hollow cone approximately 12 nm in diameter and approximately 15 nm in depth, whose axial channel tapers in width from the outside towards the inner capsid surface.  相似文献   

14.
M Yu  R H Miller  S Emerson    R H Purcell 《Journal of virology》1996,70(10):7085-7091
The capsid particle of hepadnaviruses is assembled from its dimer precursors. However, the mechanism of the protein-protein interaction is still poorly understood. A small region in the capsid protein of woodchuck hepatitis virus (WHV) contains four hydrophobic residues, including leucine 101, leucine 108, valine 115, and phenylalanine 122, that are conserved and spaced every seventh residue in the primary sequence to form a hydrophobic heptad repeat (hhr). A hydrophobic force often plays an important role in the interaction of proteins. Therefore, to investigate the role of this region in capsid assembly, we individually changed the codons specifying these four hydrophobic amino acids to codons specifying alanine or proline. In addition, we examined the in vivo infectivity of a WHV genome bearing a naturally occurring single amino acid change (histidine 104-->proline) in the hhr region. The phenotype of each altered genome was determined in both eukaryotic and prokaryotic systems by a capsid protein assay and electron microscopic examination. We show that replacement of any one of the four hydrophobic residues with alanine did not prevent capsid assembly. However, assembled capsid particles were not detected if combinations of any two of the four residues were substituted with alanines or if the spacing of these four hydrophobic residues was changed. An individual introduction of a proline (which dramatically changes the secondary structure of proteins) into different positions of this small region also abolished capsid assembly in vitro or viral replication in vivo. These results suggested that the hhr region of the core protein of WHV was critical for capsid assembly.  相似文献   

15.
The genome of a retrovirus is surrounded by a convex protein shell, or capsid, that helps facilitate infection. The major part of the capsid surface is formed by interlocking capsid protein (CA) hexamers. We report electron and X-ray crystallographic analysis of a variety of specimens assembled in vitro from Rous sarcoma virus (RSV) CA. These specimens all contain CA hexamers arranged in planar layers, modeling the authentic capsid surface. The specimens differ only in the number of layers incorporated and in the disposition of each layer with respect to its neighbor. The body of each hexamer, formed by the N-terminal domain of CA, is connected to neighboring hexamers through C-terminal domain dimerization. The resulting layer structure is very malleable due to inter-domain flexibility. A helix-capping hydrogen bond between the two domains of RSV CA creates a pivot point, which is central to controlling their relative movement. A similar mechanism for the governance of inter-domain motion was recently described for the human immunodeficiency virus type 1 (HIV-1) capsid, although there is negligible sequence identity between RSV and HIV-1 CA in the region of contact, and the amino acids involved in creating the pivot are not conserved. Our observations allow development of a physically realistic model for the way neighboring hexamers can tilt out of plane, deforming the hexamer layer and generating the continuously curved surfaces that are a feature of all retroviral capsids.  相似文献   

16.
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   

17.
The capsids of spherical viruses may contain from tens to hundreds of copies of the capsid protein(s). Despite their complexity, these particles assemble rapidly and with high fidelity. Subunit and capsid represent unique end states. However, the number of intermediate states in these reactions can be enormous-a situation analogous to the protein folding problem. Approaches to accurately model capsid assembly are still in their infancy. In this paper, we describe a sail-shaped reaction landscape, defined by the number of subunits in each species, the predicted prevalence of each species, and species stability. Prevalence can be calculated from the probability of synthesis of a given intermediate and correlates well with the appearance of intermediates in kinetics simulations. In these landscapes, we find that only those intermediates along the leading edge make a significant contribution to assembly. Although the total number of intermediates grows exponentially with capsid size, the number of leading-edge intermediates grows at a much slower rate. This result suggests that only a minute fraction of intermediates needs to be considered when describing capsid assembly.  相似文献   

18.
Of the six herpesvirus capsid proteins, the smallest capsid proteins (SCPs) share the least sequence homology among herpesvirus family members and have been implicated in virus specificity during infection. The herpes simplex virus-1 (HSV-1) SCP was shown to be horn shaped and to specifically bind the upper domain of each major capsid protein in hexons but not in pentons. In Kaposi's sarcoma-associated herpesvirus (KSHV), the protein encoded by the ORF65 gene (pORF65) is the putative SCP but its location remains controversial due to the absence of such horn-shaped densities from both the pentons and hexons of the KSHV capsid reconstructions. To directly locate the KSHV SCP, we have used electron cryomicroscopy and three-dimensional reconstruction techniques to compare the three-dimensional structure of KSHV capsids to that of anti-pORF65 antibody-labeled capsids. Our difference map shows prominent antibody densities bound to the tips of the hexons but not to pentons, indicating that KSHV SCP is attached to the upper domain of the major capsid protein in hexons but not to that in pentons, similar to HSV-1 SCP. The lack of horn-shaped densities on the hexons indicates that KSHV SCP exhibits structural features that are substantially different from those of HSV-1 SCP. The location of SCP at the outermost regions of the capsid suggests a possible role in mediating capsid interactions with the tegument and cytoskeletal proteins during infection.  相似文献   

19.
A nucleic acid-bound capsid protein dimer was previously identified using a Sindbis virus in vitro nucleocapsid assembly system and cross-linking reagents. Cross-link mapping, in combination with a model of the nucleocapsid core, suggested that this dimer contained one monomer from each of two adjacent capsomeres. This intercapsomere dimer is believed to be the initial intermediate in the nucleocapsid core assembly mechanism. This paper presents the purification of cross-linked dimers of a truncated capsid protein and the partial purification of cross-linked dimers of a full-length assembly-defective mutant. The assembly of core-like particles from these cross-linked capsid protein dimers is demonstrated. Core-like particles generated from cross-linked full-length mutant CP(19-264)L52D were examined by electron microscopy and appeared to have a morphology similar to that of wild-type in vitro-assembled core-like particles, although a slight size difference was often visible. Truncated cross-linked CP(81-264) dimers generated core-like particles as well. These core-like particles could subsequently be disassembled when reversible cross-linking reagents were used to form the dimers. The ability of the covalent intercapsomere cross-link to rescue capsid proteins with assembly defects or truncations in the amino-terminal region of the capsid protein supports the previous model of assembly and suggests a possible role for the amino-terminal region of the protein.  相似文献   

20.
Capsids of herpes simplex virus (HSV) types 1 and 2 contain seven polypeptides ranging in molecular weight from 154,000 to 12,000 (termed NC-1 through NC-7 in order of descending molecular weight). Antibodies prepared to HSV-1 capsid polypeptides isolated from sodium dodecyl sulfate-polyacrylamide gels reacted in an immunofluorescence assay against HSV-1-infected KB cells. Three of the antibodies (anti-NC-1, anti-NC-2, and anti-NC-3,4) also reacted with HSV-2-infected cells. Tryptic peptide analysis showed that each of the HSV-1 capsid polypeptides had a unique methionine peptide profile, and none appeared to be derived from the major capsid polypeptide. Comparative peptide analysis of HSV-1 and HSV-2 showed that one polypeptide (NC-7, 12,000 molecular weight) had an identical methionine peptide profile and a very similar arginine peptide profile in both virus types. The arginine peptide profile of NC-7 of HSV-1 was very different from the arginine profile of KB histone H4. Although there were certain intertypic similarities in the methionine peptide profiles of the other capsid components especially in NC-1 (the major capsid protein), there was no case where the tryptic peptides were identical in the two virus types.  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