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.
During the packaging of double-stranded DNA by bacterial viruses, the precursor procapsid loses its internal core of scaffolding protein and undergoes a substantial expansion to form the mature virion. Here we show that upon heating, purified P22 procapsids release their scaffolding protein subunits, and the coat protein lattice expands in the absence of any other cellular or viral components. Following these processes by differential scanning calorimetry revealed four different transitions that correlated with structural transitions in the coat protein shells. Exit of scaffolding protein from the procapsid occurred reversibly and just above physiological temperature. Expansion of the procapsid lattice, which was exothermic, occurred after the release of scaffolding protein. Partial denaturation of coat subunits within the intact shell structure was detected prior to the major endothermic event. This major endotherm occurred above 80 degrees C and represents particle breakage and irreversible coat protein denaturation. The results indicate that the coat subunits are designed to form a metastable precursor lattice, which appears to be separated from the mature lattice by a kinetic barrier.  相似文献   

2.
Studies on the Nucleocapsid Structure of a Group A Arbovirus   总被引:5,自引:3,他引:2       下载免费PDF全文
When Sindbis virus (273S) was treated with sodium desoxycholate, a nonhemagglutinating 136S particle was liberated from the virion, representing the viral nucleocapsid (core). Electron microscopically it appeared as a spherical particle 35 nm in diameter, showing ringlike morphological units 12 to 14 nm in diameter on its surface. When the one- and two-sided images of core particles were correlated, their structure could be demonstrated to have the T = 3 arrangement of 32 hexamer-pentamer morphological units within a symmetrical surface lattice. The core contained a further spherical structure (12 to 16 nm in diameter) which was designated as the central core component. Two proteins were found associated with the core, a third viral protein belonged to the hemagglutinating surface structures. The significance of these findings for virus classification is discussed.  相似文献   

3.
The early stages in the formation of the HIV-1 capsid (CA) protein lattice are investigated. The underlying coarse-grained (CG) model is parameterized directly from experimental data and examined under various native contact interaction strengths, CA dimer interfacial configurations, and local surface curvatures. The mechanism of early contiguous mature-style CA p6 lattice formation is explored, and a trimer-of-dimers structure is found to be crucial for CA lattice production. Quasi-equivalent generation of both the pentamer and hexamer components of the HIV-1 viral CA is also demonstrated, and the formation of pentamers is shown to be highly sensitive to local curvature, supporting the view that such inclusions in high-curvature regions allow closure of the viral CA surface. The complicated behavior of CA lattice self-assembly is shown to be reducible to a relatively simple function of the trimer-of-dimers behavior.  相似文献   

4.
Davis TL  Bonacci TM  Sprang SR  Smrcka AV 《Biochemistry》2005,44(31):10593-10604
G protein betagamma subunits associate with many binding partners in cellular signaling cascades. In previous work, we used random-peptide phage display screening to identify a diverse family of peptides that bound to a common surface on Gbetagamma subunits and blocked a subset of Gbetagamma effectors. Later studies showed that one of the peptides caused G protein activation through a novel Gbetagamma-dependent, nucleotide exchange-independent mechanism. Here we report the X-ray crystal structure of Gbeta(1)gamma(2) bound to this peptide, SIGK (SIGKAFKILGYPDYD), at 2.7 A resolution. SIGK forms a helical structure that binds the same face of Gbeta(1) as the switch II region of Galpha. The interaction interface can be subdivided into polar and nonpolar interfaces that together contain a mixture of binding determinants that may be responsible for the ability of this surface to recognize multiple protein partners. Systematic mutagenic analysis of the peptide-Gbeta(1) interface indicates that distinct sets of amino acids within this interface are required for binding of different peptides. Among these unique amino acid interactions, specific electrostatic binding contacts within the polar interface are required for peptide-mediated subunit dissociation. The data provide a mechanistic basis for multiple target recognition by Gbetagamma subunits with diverse functional interactions within a common interface and suggest that pharmacological targeting of distinct regions within this interface could allow for selective manipulation of Gbetagamma-dependent signaling pathways.  相似文献   

5.
Summary The inheritance of the high molecular weight (HMW) glutenins and of several gliadins controlled, respectively, by the long and short arms of chromosome 1B of common wheat was studied. Analysis was carried out on the progeny of two inter-varietal crosses in which the parental lines possessed differentially migrating subunits as revealed by sodium dodecyl sulphate polyacrylamide gel electrophoresis. No recombination event was detected either within the fraction of the HMW glutenins or among most of the gliadin subunits studied indicating that they are controlled by tightly linked gene clusters. One gliadin subunit (B30) showed 25.5% recombination frequency with the rest of the gliadin subunits and 23.5% recombination frequency with the fraction of the HMW glutenin subunits. It has been concluded that this subunit is controlled by a separate locus (Gld-B6), proximal to the major gliadin gene cluster on the short arm of chromosome 1B. Consequently, the recombination percentage between the glutenin loci and most of the gliadin loci was calculated as 49.0 and the distance in centi-Morgans (cM) as 53.6. The estimated distance in cM is very close to the observed recombination percentage. A genetic map of these storage protein genes is presented.  相似文献   

6.
During a 10 day-incubation on agar surfaces at 30 degrees C, cells of the gram-negative soil bacterium Pseudomonas rhodos pass through three phases distinguishable by physiological and morphological criteria. When viewed by electron microscopy, typically "rolled" mesosomes could frequently be observed in young cells. In aged cells instead, loosely rolled or stretched-out, flattened tubules could be discerned, presumed to be degenerate mesosomes. Tubular flattened structures have been isolated from these cells by lysozyme treatment or sonication and were concentrated by differential centrifugation. Electron micrographs of these preparations showed long, straight tubules which sometimes appeared sealed at one end. Their width was 34 +/- 5 nm. They contained a lining of material, which could be digested by trypsin leaving behind an electron-transparent matrix. In rare cases, isolated tubules showed a periodic fine structure composed of ellipsoidal subunits. Optical diffraction analysis yielded a lattice consisting of subunits arranged in helices of pitch-angle 27 degrees; the unit cell dimensions were shown to be 112 X 56 A. Owing to their sensitivity to trypsin, components of the regular lattice are supposed to consist of protein. It is postulated that these protein components are layered onto a tubular membrane. These tubules are clearly distinguishable by their shape and fine structure from the periodic structure of a P. rhodos cell wall layer, which exhibits a tetragonal pattern, and also from polyheads and polysheaths of defective bacteriophages. Their possible origin from intact mesosomes in discussed.  相似文献   

7.
Type IV pili are cell surface organelles found on many Gram-negative bacteria. They mediate a variety of functions, including adhesion, twitching motility, and competence for DNA uptake. The type IV pilus is a helical polymer of pilin protein subunits and is capable of rapid polymerization or depolymerization, generating large motor forces in the process. Here we show that a specific interaction between the outer membrane secretin PilQ and the type IV pilus fiber can be detected by far-Western analysis and sucrose density gradient centrifugation. Transmission electron microscopy of preparations of purified pili, to which the purified PilQ oligomer had been added, showed that PilQ was uniquely located at one end of the pilus fiber, effectively forming a "mallet-type" structure. Determination of the three-dimensional structure of the PilQ-type IV pilus complex at 26-angstroms resolution showed that the cavity within the protein complex was filled. Comparison with a previously determined structure of PilQ at 12-angstroms resolution indicated that binding of the pilus fiber induced a dissociation of the "cap" feature and lateral movement of the "arms" of the PilQ oligomer. The results demonstrate that the PilQ structure exhibits a dynamic response to the binding of its transported substrate and suggest that the secretin could play an active role in type IV pilus assembly as well as secretion.  相似文献   

8.
The S-layer lattice from Bacillus stearothermophilus DSM 2358 completely covers the cell surface and exhibits oblique symmetry. During growth of B. stearothermophilus DSM 2358 on starch medium, three amylases with molecular weights of 58,000, 98,000, and 184,000 were secreted into the culture fluid, but only the high-molecular-weight enzyme was found to be cell associated. Studies of interactions between cell wall components and amylases revealed no affinity of the high-molecular-weight amylase to isolated peptidoglycan. On the other hand, this enzyme was always found to be associated with S-layer self-assembly products or S-layer fragments released during preparation of spheroplasts by treatment of whole cells with lysozyme. The molar ratio of S-layer subunits to the bound amylase was approximately 8:1, which corresponded to one enzyme molecule per four morphological subunits. Immunoblotting experiments with polyclonal antisera against the high-molecular-weight amylase revealed a strong immunological signal in response to the enzyme but no cross-reaction with the S-layer protein or the smaller amylases. Immunogold labeling of whole cells with anti-amylase antiserum showed that the high-molecular-weight amylase is located on the outer face of the S-layer lattice. Because extraction of the amylase was possible without disintegration of the S-layer lattice into its constituent subunits, it can be excluded that the enzyme is incorporated into the crystal lattice and participates in the self-assembly process. Affinity experiments strongly suggest the presence of a specific recognition mechanism between the amylase molecules and S-layer protein domains either exposed on the outermost surface or inside the pores. In summary, results obtained in this study confirmed that the S-layer protein from B. stearothermophilus DSM 2358 functions as an adhesion site for a high-molecular-weight amylase.  相似文献   

9.
The crystal structure of satellite tobacco necrosis virus has been studied by neutron diffraction at 16 A resolution using the technique of 1H2O/2H2O solvent contrast variation to distinguish between the regions of protein and nucleic acid. The RNA density is essentially localized in a region just inside the protein coat, leading to a significant interaction between the two components. From the appearance of the RNA density we conclude that the protein coat imposes partial icosahedral symmetry on a significant proportion of the nucleic acid. The shape and dimensions of the major part of this density suggests that about 72% of the total RNA could be double-helical in structure. The most important interaction between the two components of the virus occurs between the N-terminal triple-helical arms of the protein subunits and those regions of the RNA density that could have a double-helical secondary structure.  相似文献   

10.
In this paper, the importance of charged amino and carboxyl groups for the integrity of the cell surface layer (S-layer) lattice from Bacillus coagulans E38-66 and for the self-assembly of the isolated subunits was investigated. Amidination of the free amino groups which preserved their positive net charge had no influence on both. On the other hand, acetylation and succinylation, which converted the amino groups into either neutral or negatively charged groups, and amidation of carboxyl groups were accompanied by the disintegration or at least by the loss of the regular structure of the S-layer lattice. Treatment of S-layer monolayers with the zero-length cross-linker carbodiimide led to the introduction of peptide bonds between activated carboxyl groups and amino groups from adjacent subunits. This clearly indicated that in the native S-layer lattice the charged groups are located closely enough for direct electrostatic interactions. Under disrupting conditions in which the S-layer polypeptide chains were unfolded, 58% of the Asx and Glx residues could be amidated, indicating that they occur in the free carboxylic acid form. As derived from chemical modification of monolayer self-assembly products, about 90% of the lysine and 70% of the aspartic and glutamic acid residues are aligned on the surface of the S-layer protein domains. This corresponded to 45 amino groups and to 63 carboxyl groups per S-layer subunit. Labelling experiments with macromolecules with different sizes and charges and adsorption studies with ion-exchange particles revealed a surplus of free carboxyl groups on the inner and on the outer faces of the S-layer lattice. Since the carboxyl groups on the outer S-layer face were accessible only for protein molecules significantly smaller then the S-layer protomers or for positively charged, thin polymer chains extending from the surface of ion-exchange beads, the negatively charged sites must be located within indentations of the corrugated S-layer protein network. This was in contrast to the carboxyl groups on the inner S-layer face, which were found to be exposed on elevations of the S-layer protein domains (D. Pum, M. Sára, and U.B. Sleytr, J. Bacteriol. 171:5296-5303, 1989).  相似文献   

11.
Assembly of certain classes of bacterial and animal viruses requires the transient presence of molecules known as scaffolding proteins, which are essential for the assembly of the precursor procapsid. To assemble a procapsid of the proper size, each viral coat subunit must adopt the correct quasiequivalent conformation from several possible choices, depending upon the T number of the capsid. In the absence of scaffolding protein, the viral coat proteins form aberrantly shaped and incorrectly sized capsids that cannot package DNA. Although scaffolding proteins do not form icosahedral cores within procapsids, an icosahedrally ordered coat/scaffolding interaction could explain how scaffolding can cause conformational differences between coat subunits. To identify the interaction sites of scaffolding protein with the bacteriophage P22 coat protein lattice, we have determined electron cryomicroscopy structures of scaffolding-containing and scaffolding-lacking procapsids. The resulting difference maps suggest specific interactions of scaffolding protein with only four of the seven quasiequivalent coat protein conformations in the T = 7 P22 procapsid lattice, supporting the idea that the conformational switching of a coat subunit is regulated by the type of interactions it undergoes with the scaffolding protein. Based on these results, we propose a model for P22 procapsid assembly that involves alternating steps in which first coat, then scaffolding subunits form self-interactions that promote the addition of the other protein. Together, the coat and scaffolding provide overlapping sets of binding interactions that drive the formation of the procapsid.  相似文献   

12.
During a 10 day-incubation on agar surfaces at 30°C, cells of the gram-negative soil bacterium Pseudomonas rhodos pass through three phases distinguishable by physiological and morphological criteria. When viewed by electron microscopy, typically “rolled” mesosomes could frequently be observed in young cells. In aged cells instead, loosely rolled or stretched-out, flattened tubules could be discerned, presumed to be degenerate mesosomes. Tubular flattened structures have been isolated from these cells by lysozyme treatment or sonication and were concentrated by differential centrifugation. Electron micrographs of these preparations showed long, straight tubules which sometimes appeared sealed at one end. Their width was 34±5 nm. They contained a lining of material, which could be digested by trypsin leaving behind an electron-transparent matrix. In rare cases, isolated tubules showed a periodic fine structure composed of ellipsoidal subunits. Optical diffraction analysis yielded a lattice consisting of subunits arranged in helices of pitch-angle 27°; the unit cell dimensions were shown to be 112×56 Å. Owing to their sensitivity to trypsin, components of the regular lattice are supposed to consist of protein. It is postulated that these protein components are layered onto a tubular membrane. These tubules are clearly distinguishable by their shape and fine structure from the periodic structure of a P. rhodos cell wall layer, which exhibits a tetragonal pattern, and also from polyheads and polysheaths of defective bacteriophages. Their possible origin from intact mesosomes is discussed.  相似文献   

13.
The formation of closed icosahedral capsids from a single species of coat protein subunit requires that the subunits assume different conformations at different lattice positions. In the double-stranded DNA bacteriophage P22, formation of correctly dimensioned capsids is mediated by interaction between coat protein subunits and scaffolding protein. Raman spectroscopy has been employed to compare the conformations of coat protein subunits which have been polymerized to form capsids in the presence and absence of the of scaffolding protein display a Raman spectrum characterized by a broad amide I band centered at 1665 cm-1 with a discernible shoulder near 1653 cm-1, and a broad amide III profile centered at 1238 cm-1 but asymmetrically skewed to higher frequency. These spectral features indicate that the protein conformation in procapsid shells is rich in beta-sheet secondary structure but contains also a significant distribution of alpha-helix. When biologically active, purified subunits assemble in the absence of scaffolding protein, they form polydisperse multimers lacking the proper dimensions of procapsid closed shells. We designate these multimers as "associated subunits" (AS). The Raman spectrum of associated subunits indicates a narrower distribution of secondary structure. The associated subunits are characterized by a sharper and more intense Raman amide I band at 1666 cm-1, with no prominent amide I shoulder of lower frequency. An analogous narrowing of the Raman amide III profile is also observed for AS particles, with an accompanying shift of the amide III band center to 1235 cm-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The coat and scaffolding proteins of bacteriophage P22 procapsids have been purified in soluble form. By incubating both purified proteins with a mutant-infected cell extract lacking procapsids, but competent for DNA packaging in vitro (Poteete et al., 1979), we were able to obtain assembly of biologically active procapsids in vitro. The active species for complementation in vitro in both protein preparations copurified with the soluble subunits, indicating that these subunits represent precursors in procapsid polymerization.When the purified coat and scaffolding subunits were mixed directly, they polymerized into double-shelled procapsid-like structures during dialysis from 1.5 m-guanidine hydrochloride to buffer. When dialyzed separately under the same conditions, the scaffolding subunits did not polymerize but remained as soluble subunits, as did most of the coat subunits. No evidence was found for self-assembly of the scaffolding protein in the absence of the coat protein.The unassembled coat subunits sedimented at 3.9 S and the unassembled scaffolding subunits sedimented at 2.4 S in sucrose gradients. The Stokes' radius, determined by gel filtration, was 25 Å for the coat subunits and 34 Å for the scaffolding subunits. These results indicate that the scaffolding subunits are relatively slender elongated molecules, whereas the coat subunits are more globular.The experiments suggest that the procapsid is built by copolymerization of the two protein species. Their interaction on the growing surface of the shell structure, and not in solution, appears to regulate successive binding interactions.  相似文献   

15.
Tau protein plays a role in the extension and maintenance of neuronal processes through a direct association with microtubules. To characterize the nature of this association, we have synthesized a collection of tau protein fragments and studied their binding properties. The relatively weak affinity of tau protein for microtubules (approximately 10(-7) M) is concentrated in a large region containing three or four 18 amino acid repeated binding elements. These are separated by apparently flexible but less conserved linker sequences of 13-14 amino acids that do not bind. Within the repeats, the binding energy for microtubules is delocalized and derives from a series of weak interactions contributed by small groups of amino acids. These unusual characteristics suggest tau protein can assume multiple conformations and can pivot and perhaps migrate on the surface of the microtubule. The flexible structure of the tau protein binding interaction may allow it to be easily displaced from the microtubule lattice and may have important consequences for its function.  相似文献   

16.
The outer dynein arms of Chlamydomonas flagella are attached to a precise site on the outer doublet microtubules and repeat at a regular interval of 24 nm. This binding is mediated by the outer dynein arm docking complex (ODA-DC), which is composed of three protein subunits. In this study, antibodies against the 83- and 62-kD subunits (DC83 and DC62) of the ODA-DC were used to analyze its state of association with outer arm components within the cytoplasm, and its localization in the axonemes of oda mutants. Immunoprecipitation indicates that DC83 and DC62 are preassembled within the cytoplasm, but that they are not associated with outer arm dynein. Both proteins are lost or greatly diminished in oda1 and oda3, mutants in the structural genes of DC62 and DC83, respectively, demonstrating that their association is necessary for their stable presence in the cytoplasm. Immunoelectron microscopy indicates that DC83 repeats at 24-nm intervals along the length of the doublet microtubules of oda6, which lacks outer arms; thus, outer arm periodicity may be determined by the ODA-DC. Flagellar regeneration and temporary dikaryon experiments indicate that the ODA-DC can be rapidly transported into the flagellum and assembled on the doublet microtubules independently of the outer arms and independently of flagellar growth. Unexpectedly, the intensity of ODA-DC labeling decreased toward the distal ends of axonemes of oda6 but not wild-type cells, suggesting that the outer arms reciprocally contribute to the assembly/stability of the ODA-DC.  相似文献   

17.
Abstract: Fimbriae are long filamentous polymeric protein structures located at the surface of bacterial cells. They enable the bacteria to bind to specific receptor structures and thereby to colonise specific surfaces. Fimbriae consist of so-called major and minor subunits, which form, in a specific order, the fimbrial structure. In this review emphasis is put on the genetic organisation, regulation and especially on the biosynthesis of fimbriae of enterotoxigenic Escherichia coli strains, and more in particular on K88 and related fimbriae, with ample reference to the well-studied P and type 1 fimbriae. The biosynthesis of these fimbriae requires two specific and unique proteins, a periplasmic chaperone and an outer membrane located molecular usher ('doorkeeper'). Molecular and structural aspects of the secretion of fimbrial subunits across the cytoplasmic membrane, the interaction of these subunits with the periplasmic molecular chaperone, their translocation to the inner site of the outer membrane and their interaction with the usher protein, as well as the (ordered) translocation of the subunits across the outer membrane and their assembly into a grwoing fimbrial structure will be described. A model for K88 fimbriae is presented.  相似文献   

18.
Two related forms of the respiratory chain NADH dehydrogenase (NADH:ubiquinone reductase or complex I) are synthesized in the mitochondria of Neurospora crassa. Normally growing cells make a large form that consists of 25 subunits encoded by nuclear DNA and six to seven subunits encoded by mitochondrial DNA. Cells grown in the presence of chloramphenicol, however, make a smaller form comprising only 13 subunits, all encoded by nuclear DNA. When the large enzyme is dissected by chaotropic agents (such as NaBr), all those subunits of the large form that are missing in the small form can be isolated as a distinct, so-called hydrophobic fragment. The small enzyme and the hydrophobic fragment make up, with regard to their redox groups, subunit composition and function, two complementary parts of the large-form NADH dehydrogenase. Averaging of electron microscope images of single particles of the large enzyme was carried out, revealing an unusual L-shaped structure with two domains or "arms" arranged at right angles. The hydrophobic fragment obtained by the NaBr treatment corresponds in size and appearance to one of these arms. A three-dimensional reconstruction from images of negatively stained membrane crystals of the large-form NADH dehydrogenase shows a peripheral domain, protruding from the membrane, with weak unresolved density within the membrane. This peripheral domain was removed by washing the crystals in situ with 2 M-NaBr, exposing a large membrane-buried domain, which was reconstructed in three dimensions. A three-dimensional reconstruction of the small enzyme from negatively stained membrane crystals, also described here, shows only a peripheral domain. These results suggest that the membrane protruding arm of the large form corresponds to the small enzyme, whereas the arm lying within the membrane can be identified as the hydrophobic fragment. The two parts of NADH dehydrogenase that can be defined by the separate genetic origin of (most of) their subunits, their independent assembly, and their distinct contributions to the electron pathway can thus be assigned to the two arms of the L-shaped complex I.  相似文献   

19.
The sulphur-dependent archaebacterium Thermoproteus tenax has a cylindrical cell shape variable in length, but constant in diameter. Its whole surface is covered by a regular protein layer (S-layer). The lattice has p6 symmetry and a lattice constant of 32.8 nm. The three-dimensional reconstruction from a tilt series of isolated and negatively stained S-layer shows a complex mass distribution of the protein: a prominent, pillar-shaped protrusion is located at the 6-fold crystallographic axis with radiating arms connecting neighbouring hexamers in the vicinity of the 3-fold axis. The base vectors of the S-layer lattice have a preferred orientation with respect to the longitudinal axis of the cell. The layer can be seen as a helical structure consisting of a right-handed, two-stranded helix, with the individual chains running parallel. Supposing that new S-layer protein is inserted at lattice faults (wedge disclinations) near the poles, growing of the layer would then proceed by moving a disclination at the end of the helix. The constant shape of the cell, as well as the particular structure of the layer, strongly suggest that this S-layer has a shape-maintaining function.  相似文献   

20.
The biogenesis of the cytoskeletal proteins actin and tubulin involves interaction of nascent chains of each of the two proteins with the oligomeric protein prefoldin (PFD) and their subsequent transfer to the cytosolic chaperonin CCT (chaperonin containing TCP-1). Here we show by electron microscopy that eukaryotic PFD, which has a similar structure to its archaeal counterpart, interacts with unfolded actin along the tips of its projecting arms. In its PFD-bound state, actin seems to acquire a conformation similar to that adopted when it is bound to CCT. Three-dimensional reconstruction of the CCT:PFD complex based on cryoelectron microscopy reveals that PFD binds to each of the CCT rings in a unique conformation through two specific CCT subunits that are placed in a 1,4 arrangement. This defines the phasing of the CCT rings and suggests a handoff mechanism for PFD.  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