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 9 毫秒
1.
Sesbania mosaic virus particles consist of 180 coat protein subunits of 29kDa organized on a T=3 icosahedral lattice. N-terminal deletion mutants of coat protein that lack 36 (CP-NDelta36) and 65 (CP-NDelta65) residues from the N terminus, when expressed in Escherichia coli, produced similar T=1 capsids of approximate diameter 20nm. In contrast to the wild-type particles, these contain only 60 copies of the truncated protein subunits (T=1). CP-NDelta65 lacks the "beta-annulus" believed to be responsible for the error-free assembly of T=3 particles. Though the CP-NDelta36 mutant has the beta-annulus segment, it does not form a T=3 capsid, presumably because it lacks an arginine-rich motif found close to the amino terminus. Both CP-NDelta36 and CP-NDelta65 T=1 capsids retain many key features of the T=3 quaternary structure. Calcium binding geometries at the coat protein interfaces in these two particles are also nearly identical. When the conserved aspartate residues that coordinate the calcium, D146 and D149 in the CP-NDelta65, were mutated to asparagine (CP-NDelta65-D146N-D149N), the subunits assembled into T=1 particles but failed to bind calcium ions. The structure of this mutant revealed particles that were slightly expanded. The analysis of the structures of these mutant capsids suggests that although calcium binding contributes substantially to the stability of T=1 particles, it is not mandatory for their assembly. In contrast, the presence of a large fraction of the amino-terminal arm including sequences that precede the beta-annulus and the conserved D149 appear to be indispensable for the error-free assembly of T=3 particles.  相似文献   

2.
3.
4.
Cucumber necrosis virus (CNV) is a member of the genus Tombusvirus, of which tomato bushy stunt virus (TBSV) is the type member. The capsid protein for this group of viruses is composed of three major domains: the R domain, which interacts with the RNA genome: the S domain, which forms the tight capsid shell: and the protruding P domain, which extends approximately 40 Angstrom from the surface. Here, we present the cryo-transmission electron microscopy structures of both the T=1 and T=3 capsids to a resolution of approximately 12 Angstrom. The T=3 capsid is essentially identical with that of TBSV, and the T=1 particles are well described by the A subunit pentons from TBSV. Perhaps most notable is the fact that the T=3 particles have an articulated internal structure with two major internal shells, while the internal core of the T=1 particle is essentially disordered. These internal shells of the T=3 capsid agree extremely well in both dimension and character with published neutron-scattering results. This structure, combined with mutagenesis results in the accompanying article, suggests that the R domain forms an internal icosahedral scaffold that may play a role in T=3 capsid assembly. In addition, the N-terminal region has been shown to be involved in chloroplast targeting. Therefore, this region apparently has remarkably diverse functions that may be distributed unevenly among the quasi-equivalent A, B, and C subunits.  相似文献   

5.
Reassembled alfalfa mosaic virus coat protein was partially digested with trypsin to remove the first 26 amino acids (Bol et al., 1974). These particles are empty icosahedral protein shells built with 60 alfalfa mosaic virus protein subunits. This aggregate has been crystallized in two different crystal forms, one of which diffracts X-rays to at least 3.4 Å resolution. The type I crystals (space group P63, a = 200 A?, c = 314 A?) contain two particles per cell separated by 195 Å with each sitting on a 3-fold axis. The type II crystals contain three particles per cell in space group P31or P32 (a = 201 A?, c = 485 A?). Other T = 1 viral particles have very similar diameters.  相似文献   

6.
Sesbania mosaic virus (SeMV) capsids are stabilized by RNA-protein, protein-protein and calcium-mediated protein-protein interactions. The removal of calcium has been proposed to be a prerequisite for the disassembly of the virus. The crystal structure of native T=3 SeMV capsid revealed that residues D146 and D149 from one subunit and Y205, N267 and N268 of the neighboring subunit form the calcium-binding site (CBS). The CBS environment is found to be identical even in the recombinant CP-NDelta65 T=1 capsids. Here, we have addressed the role of calcium and the residues involved in calcium co-ordination in the assembly and stability of T=3 and T=1 capsids by mutational analysis. Deletion of N267 and N268 did not affect T=3 or T=1 assembly, although the capsids were devoid of calcium, suggesting that assembly does not require calcium ions. However, the stability of the capsids was reduced drastically. Site-directed mutagenesis revealed that either a single mutation (D149N) or a double mutation (D146N-D149N) of SeMV coat protein affected drastically both the assembly and stability of T=3 capsids. On the other hand, the D146N-D149N mutation in CP-NDelta65 did not affect the assembly of T=1 capsid, although their stability was reduced considerably. Since the major difference between the T=3 and T=1 capsids is the absence of the N-terminal arginine-rich motif (N-ARM) and the beta-annulus from the subunits forming the T=1 capsids, it is possible that D149 initiates the N-ARM-RNA interactions that lead to the formation of the beta-annulus, which is essential for T=3 capsid assembly.  相似文献   

7.
The vast majority of viruses consist of a nucleic acid surrounded by a protective icosahedral protein shell called the capsid. During viral infection of a host cell, the timing and efficiency of the assembly process is important for ensuring the production of infectious new progeny virus particles. In the class of single-stranded RNA (ssRNA) viruses, the assembly of the capsid takes place in tandem with packaging of the ssRNA genome in a highly cooperative co-assembly process. In simple ssRNA viruses such as the bacteriophage MS2 and small RNA plant viruses such as STNV, this cooperative process results from multiple interactions between the protein shell and sites in the RNA genome which have been termed packaging signals. Using a stochastic assembly algorithm which includes cooperative interactions between the protein shell and packaging signals in the RNA genome, we demonstrate that highly efficient assembly of STNV capsids arises from a set of simple local rules. Altering the local assembly rules results in different nucleation scenarios with varying assembly efficiencies, which in some cases depend strongly on interactions with RNA packaging signals. Our results provide a potential simple explanation based on local assembly rules for the ability of some ssRNA viruses to spontaneously assemble around charged polymers and other non-viral RNAs in vitro.  相似文献   

8.
S D Fuller 《Cell》1987,48(6):923-934
The three-dimensional structure of Sindbis virus, an enveloped animal virus, has been determined to a resolution of 35 A by using a common lines procedure to combine cryoelectron micrographs of vitrified particles. The spikes of the virus appear as columnar trimers arranged on a T=4 lattice. The lipid bilayer of the virus envelope is polyhedral and surrounds a smooth T=3 nucleocapsid. Hence, a complete Sindbis virion (molecular weight 46.4 X 10(6)) contains 240 copies of each of the spike proteins and 180 copies of the capsid protein. The arrangement of the spike proteins is complementary to that of the nucleocapsid. Two types of spike-capsid interactions are seen. Spike trimers near the fivefold axes interact tightly with triplets of capsid elements, whereas those on the threefold axes interact more loosely.  相似文献   

9.
K. Fukuyama, S. S. Abdel-Meguid, J. E. Johnson, and M. G. Rossmann (J. Mol. Biol. 167:873-984, 1983) reported the structure of alfalfa mosaic virus assembled from the capsid protein as a T=1 icosahedral empty particle at 4.5-A resolution. The information contained in the structure included the particle size, protein shell thickness, presence of wide holes at the icosahedral fivefold axes, and a proposal that the capsid protein adopts a beta-barrel structure. In the present work, the X-ray diffraction data of Fukuyama et al. as well as the data subsequently collected by I. Fita, Y. Hata, and M. G. Rossmann (unpublished) were reprocessed to 4.0-A resolution, and the structure was solved by molecular replacement. The current structure allowed the tracing of the polypeptide chain of the capsid protein confirming the beta-sandwich fold and provides information on intersubunit interactions in the particle. However, it was not possible to definitively assign the amino acid sequence to the side chain density at 4-A resolution. The particle structure was also determined by cryoelectron microscopy and image reconstruction methods and found to be in excellent agreement with the X-ray model.  相似文献   

10.
GAPDH gene diversity in spirochetes: a paradigm for genetic promiscuity.   总被引:4,自引:0,他引:4  
In this study we have determined gap sequences from nine different spirochetes. Phylogenetic analyses of these sequences in the context of all other available eubacterial and a selection of eukaryotic Gap sequences demonstrated that the eubacterial glyceraldehyde-3-phosphate dehydrogenase (GAPDH) gene diversity encompasses at least five highly distinct gene families. Within these gene families, spirochetes show an extreme degree of sequence divergence that is probably the result of several lateral gene transfer events between spirochetes and other eubacterial phyla, and early gene duplications in the eubacterial ancestor. A Gap1 sequence from the syphilis spirochete Treponema pallidum has recently been shown to be closely related to GapC sequences from Euglenozoa. Here we demonstrate that several other spirochetal species are part of this cluster, supporting the conclusion that an interkingdom gene transfer from spirochetes to Euglenozoa must have occurred. Furthermore, we provide evidence that the GAPDH genes present in the protists Parabasalia may also be of spirochetal descent.  相似文献   

11.
12.
A T = 1 empty aggregate of alfalfa mosaic virus coat protein had been crystallized in a hexagonal unit cell and its orientation was determined with the rotation function. A single heavy-atom derivative has now been prepared and the position of the two Hg atoms per protein subunit were determined using a systematic Patterson search procedure, given the particle orientation. Phases, initially determined by single isomorphous replacement, were refined by six cycles of electron density averaging and solvent leveling to produce a 4.5 A resolution electron density map. The protein coat is confined between 95 and 58 A radius. The subunit boundary could be delineated easily. It has a central cavity reminiscent of the beta-barrel in other spherical plant viruses, but its topology could not be determined unambiguously. The spherical particle has large holes at the 5-fold axes, consistent with previous observations. The subunits have substantial interactions at the 2 and 3-fold axes. The structure of the elongated particles is discussed in relation to these results.  相似文献   

13.
In polyomaviruses the pentameric capsomers are interlinked by the long C-terminal arm of the structural protein VP1. The T=7 icosahedral structure of these viruses is possible due to an intriguing adaptability of this linker arm to the different local environments in the capsid. To explore the assembly process, we have compared the structure of two virus-like particles (VLPs) formed, as we found, in a calcium-dependent manner by the VP1 protein of human polyomavirus BK. The structures were determined using electron cryomicroscopy (cryo-EM), and the three-dimensional reconstructions were interpreted by atomic modeling. In the small VP1 particle, 26.4 nm in diameter, the pentameric capsomers form an icosahedral T=1 surface lattice with meeting densities at the threefold axes that interlinked three capsomers. In the larger particle, 50.6 nm in diameter, the capsomers form a T=7 icosahedral shell with three unique contacts. A folding model of the BKV VP1 protein was obtained by alignment with the VP1 protein of simian virus 40 (SV40). The model fitted well into the cryo-EM density of the T=7 particle. However, residues 297 to 362 of the C-terminal arm had to be remodeled to accommodate the higher curvature of the T=1 particle. The loops, before and after the C-terminal short helix, were shown to provide the hinges that allowed curvature variation in the particle shell. The meeting densities seen at the threefold axes in the T=1 particle were consistent with the triple-helix interlinking contact at the local threefold axes in the T=7 structure.  相似文献   

14.
Gammadelta T cells remain an enigma. They are capable of generating more unique antigen receptors than alphabeta T cells and B cells combined, yet their repertoire of antigen receptors is dominated by specific subsets that recognize a limited number of antigens. A variety of sometimes conflicting effector functions have been ascribed to them, yet their biological function(s) remains unclear. On the basis of studies of gammadelta T cells in infectious and autoimmune diseases, we argue that gammadelta T cells perform different functions according to their tissue distribution, antigen-receptor structure and local microenvironment; we also discuss how and at what stage of the immune response they become activated.  相似文献   

15.
E F Wagner  B Auer    M Schweiger 《Journal of virology》1979,29(3):1229-1231
The bacterial virus T1 grows interchangeably on different Escherichia coli strains (C, B, and K). This implies that T1 has an efficient mechanism to overcome the host restriction barrier. The DNA of T1 was found to be methylated independently of the hosts. The percentage of N6-methyladenine varied from 1.6 to 1.8, and the 5-methylcytosine content varied from 0.1 to 0.4%. In contrast, the range in percentage of N6-methyladenine and 5-methylcytosine found in the hosts was 0.7 to 2.4 and 0.0 to 1.1, respectively.  相似文献   

16.
17.
G Li  M Simm  M J Potash    D J Volsky 《Journal of virology》1993,67(7):3969-3977
Human immunodeficiency virus type 1 (HIV-1) replicates efficiently in nonproliferating monocytes and macrophages but not in resting primary T lymphocytes. To determine the contribution of cell division to the HIV-1 replicative cycle in T cells, we evaluated HIV-1 expression, integration of proviral DNA, and production of infectious progeny virus in C8166 T-lymphoid cells blocked in cell division by treatment with either mitomycin, a DNA cross-linker, or aphidicolin, a DNA polymerase alpha inhibitor. The arrest of cell division was confirmed by assay of [3H]thymidine uptake; the nondividing cells remained viable for at least 3 days after treatment. HIV-1 was expressed and replicated equally well in nondividing and dividing C8166 cells, as judged by the comparison of the levels of p24 core antigens in culture supernatants, the proportion of cells expressing HIV-1 specific antigens, the pattern and quantity of HIV-1 DNA present in the extrachromosomal and total cellular DNA fractions, and the biological activity of progeny viruses. A polymerase chain reaction-based viral DNA integration assay indicated that HIV-1 provirus was integrated in C8166 cells treated with either of the two inhibitors of cell division. Similar results were obtained by using growth-arrested Jurkat T-lymphoid cells. We conclude that cell division and cellular DNA synthesis are not required for efficient HIV-1 expression in T cells.  相似文献   

18.
Nudaurelia capensis ω virus-like particles have been characterized as a 480-Å procapsid and a 410-Å capsid, both with T=4 quasisymmetry. Procapsids transition to capsids when pH is lowered from 7.6 to 5.0. Capsids undergo autoproteolysis at residue 570, generating the 74-residue C-terminal polypeptide that remains with the particle. Here we show that the particle size becomes smaller under conditions between pH 6.8 and 6.0 without activating cleavage and that the particle remains at an intermediate size when the pH is carefully maintained. At pH 5.8, cleavage is very slow, becoming detectable only after 9 h. The optimum pH for cleavage is 5.0 (half-life, ~30 min), with a significant reduction in the cleavage rate at pH values below 5. We also show that lowering the pH is required only to make the virus particles compact and to presumably form the active site for autoproteolysis but not for the chemistry of cleavage. The cleavage reaction proceeds at pH 7.0 after ~10% of the subunits cleave at pH 5.0. Employing the virion crystal structure for reference, we investigated the role of electrostatic repulsion of acidic residues in the pH-dependent large conformational changes. Three mutations of Glu to Gln that formed procapsids showed three different phenotypes on maturation. One, close to the threefold and quasithreefold symmetry axes and far from the cleavage site, did not mature at pH 5, and electron cryomicroscopy reconstruction showed that it was intermediate in size between those of the procapsid and capsid; one near the cleavage site exhibited a wild-type phenotype; and a third, far from the cleavage site, resulted in cleavage of 50% of the subunits after 4 h, suggesting quasiequivalent specificity of the mutation.  相似文献   

19.
Summary T1 infected bacteria exhibit a distinct pattern of gene expression. The control of this expression is accessible to biochemical analysis. T1 induces the synthesis of 31 proteins inE. coli. The virion contains 15 proteins. By means of T1 amber mutants, 10 gene products have been assigned to specific T1 genes. Three classes of T1 proteins are defined by the kinetics of their syntheses: early, early-late and late proteins. The regulation of protein synthesis involves at least three mechanisms: for cessation of host gene expression, for discontinuation of the early class during the late phase and for induction of the late T1 proteins. The positive control of late gene expression is not coupled to replication. The host RNA-polymerase transcribes the viral genome throughout the infectious cycle. No virus coded RNA-polymerase is induced.  相似文献   

20.
Accumulating evidence indicates that plant disease-resistance (R) proteins assemble in hetero-multimeric protein complexes in the absence of pathogens. Such complexes might enable the indirect recognition of pathogen effector molecules during attempted pathogen invasion. RAR1 and SGT1 are required for the function of most known R proteins. They interact with each other and with diverse protein complexes, which might explain their multi-functionality. The promiscuous behavior of RAR1 and SGT1 might be crucial for the formation and activation of R protein-containing recognition complexes as well as for regulating downstream signaling processes.  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