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 0 毫秒
1.
2.
3.
We have determined the crystal structure of the enzyme-product complex of the hammerhead ribozyme by using a reinforced crystal lattice to trap the complex prior to dissociation and by employing X-ray holographic image reconstruction, a real-space electron density imaging and refinement procedure. Subsequent to catalysis, the cleavage site residue (C-17), together with its 2',3'-cyclic phosphate, adopts a conformation close to and approximately perpendicular to the Watson-Crick base-pairing faces of two highly conserved purines in the ribozyme's catalytic pocket (G-5 and A-6). We observe several interactions with functional groups on these residues that have been identified as critical for ribozyme activity by biochemical analyses but whose role has defied explanation in terms of previous structural analyses. These interactions may therefore be relevant to the hammerhead ribozyme reaction mechanism.  相似文献   

4.
5.
6.
Electron crystallography determines the structure of membrane proteins and other periodic samples by recording either images or diffraction patterns. Computer processing of recorded images requires the determination of the reciprocal lattice parameters in the Fourier transform of the image. We have developed a set of three programs 2dx_peaksearch, 2dx_findlat and 2dx_getlat, which can determine the reciprocal lattice from a Fourier transformation of a 2D crystal image automatically. 2dx_peaksearch determines a list of Fourier peak coordinates from a processed calculated diffraction pattern. These coordinates are evaluated by 2dx_findlat to determine one or more lattices, using a-priori knowledge of the real-space crystal unit cell dimensions, and the sample tilt geometry. If these are unknown, then the program 2dx_getlat can be used to obtain a guess for the unit cell dimensions. These programs are available as part of the 2dx software package for the image processing of 2D crystal images at http://2dx.org.  相似文献   

7.
Catalytic activities can be facilitated by ordered enzymatic arrays that co-localize and orient enzymes and their substrates. The purified RNA-dependent RNA polymerase from poliovirus self-assembles to form two-dimensional lattices, possibly facilitating the assembly of viral RNA replication complexes on the cytoplasmic face of intracellular membranes. Creation of a two-dimensional lattice requires at least two different molecular contacts between polymerase molecules. One set of polymerase contacts, between the “thumb” domain of one polymerase and the back of the “palm” domain of another, has been previously defined. To identify the second interface needed for lattice formation and to test its function in viral RNA synthesis, we used a hybrid approach of electron microscopic and biochemical evaluation of both wild-type and mutant viral polymerases to evaluate computationally generated models of this second interface. A unique solution satisfied all constraints and predicted a two-dimensional structure formed from antiparallel arrays of polymerase fibers that use contacts from the flexible amino-terminal region of the protein. Enzymes that contained mutations in this newly defined interface did not form lattices and altered the structure of wild-type lattices. When reconstructed into virus, mutations that disrupt lattice assembly exhibited growth defects, synthetic lethality or both, supporting the function of the oligomeric lattice in infected cells. Understanding the structure of polymerase lattices within the multimeric RNA-dependent RNA polymerase complex should facilitate antiviral drug design and provide a precedent for other positive-strand RNA viruses.  相似文献   

8.
Erysimum latent virus, a tymovirus, contains 180 protein subunits arranged in a T = 3 icosahedral surface lattice. A cubic crystal form (with space group P213 and a = 414 A?) and a monoclinic form (space group B2, a = 442 A?, b = 422 A?, c = 387 A?, γ = 95 °) have been observed. The asymmetric units of the two crystal forms contain one-third and one whole virus particle, respectively. Two possible packing arrangements of the virus particles in the monoclinic unit cell have been deduced from the low-angle diffraction patterns. X-ray diffraction data from the monoclinic crystals extend to at least 3·7 Å resolution.  相似文献   

9.
Non-crystallographic symmetry in the crystal dimer of wheat germ agglutinin   总被引:1,自引:0,他引:1  
Three isomorphous heavy atom derivatives of wheat germ agglutinin crystals, KAu(CN)2, K2Pt(NH3)2(NO)2 and mersalyl, have been examined at high resolution. Heavy atom sites were located from difference Patterson maps in three dimensions at 2.15 Å resolution for the gold and platinum derivatives and with less certainty in the centrosymmetric [010] projection for the mersalyl derivative. These sites are distributed in the crystallographic asymmetric unit such that one half of them can be related to the other half by a 180 ° rotation about an axis parallel to a, and an additional translation of about 6.35 Å along that axis. It is suggested that the two subunits of the wheat germ agglutinin dimer, which represent the asymmetric unit of the C2 unit cell, are related by the same symmetry axis, causing heterologous subunit contacts due to the 6.35 Å translation of one relative to the other subunit.  相似文献   

10.
11.
12.
After recognizing and binding to its host cell, poliovirus (like other nonenveloped viruses) faces the challenge of translocating its genome across a cellular membrane and into the cytoplasm. To avoid entanglement with the capsid, the RNA must exit via a single site on the virion surface. However, the mechanism by which a single site is selected (from among 60 equivalents) is unknown; and until now, even its location on the virion surface has been controversial. To help to elucidate the mechanism of infection, we have used single-particle cryo-electron microscopy and tomography to reconstruct conformationally altered intermediates that are formed by the poliovirion at various stages of the poliovirus infection process. Recently, we reported icosahedrally symmetric structures for two forms of the end-state 80S empty capsid particle. Surprisingly, RNA was frequently visible near the capsid; and in a subset of the virions, RNA was seen on both the inside and outside of the capsid, caught in the act of exiting. To visualize RNA exiting, we have now determined asymmetric reconstructions from that subset, using both single-particle cryo-electron microscopy and cryo-electron tomographic methods, producing independent reconstructions at ∼50-Å resolution. Contrary to predictions in the literature, the footprint of RNA on the capsid surface is located close to a viral 2-fold axis, covering a slot-shaped area of reduced density that is present in both of the symmetrized 80S reconstructions and which extends by about 20 Å away from the 2-fold axis toward each neighboring 5-fold axis.In its role as the intermediate that links one round of infection with the next, a virus particle protects the viral genome during passage from cell to cell and from host to host, it specifically recognizes and binds to target cells, and it delivers the viral genome into the appropriate compartment in the target cell. For enveloped viruses, which have their own external membranes, fusion of the viral membrane with a host membrane presents a conceptually simple mechanism for delivery of the genome or nucleoprotein into the cytoplasm. For nonenveloped viruses, the viral particle must provide the machinery necessary for either the entire virion, a nucleoprotein complex, or the viral genome to cross a membrane. This process remains poorly understood. Poliovirus provides an excellent model system for probing the mechanisms used for genome translocation. As the type member of the Picornavirus family and the etiological agent of poliomyelitis, poliovirus has been well characterized biochemically and genetically (42), its cell entry pathways have been well characterized (5, 15, 30, 52), and a number of cell entry intermediates have been identified and are accessible for structural studies (2-4, 7, 8, 18, 34, 38, 42, 55, 56).The capsid of the mature poliovirion (160S particle) consists of 60 copies of each of the four coat proteins VP1, VP2, VP3, and VP4 (which is myristolated at its amino-terminal glycine [13]) and encloses a 7.5-kbp positive-sense RNA genome. The outer surface of the capsid has a number of major features, including star-shaped mesas at its 5-fold axes, 3-fold propeller-like protrusions, canyon-like depressions surrounding each of the 5-fold mesas, and depressions at the 2-fold axes (30, 31).Poliovirus infection is initiated when the virus binds to the host-cell-surface poliovirus receptor (called Pvr or CD155) (41), triggering a conformational change of the native capsid into an altered particle called the A particle or 135S particle (18, 19). The 135S particle has been shown to be expanded by about 4% (2, 7), is infectious (16, 33), and is believed to be a productive intermediate in viral entry (30, 33). This conformational change results in the externalization of the small myristoylated capsid protein, VP4 (18), and of the amino-terminal extension of VP1 (which includes a conserved amphipathic helix) (23). Both of these externalized polypeptides then associate with membranes (17, 23). In subsequent steps, the viral genome is released from the capsid and translocated across a membrane (probably an endosomal membrane [5]) to gain access to the cytoplasm, leaving behind an end-state empty capsid shell (called the 80S particle). The trigger for RNA release and the mechanism of genome translocation are both poorly understood (30, 52).Electrophysiology and mutational experiments have shown that the externalization of VP4 and of the amino terminus of VP1 is associated with the formation of channels in membranes (17, 49, 50) and, furthermore, that point mutations in threonine 28 of VP4 can either eliminate (T28G) or alter (T28V, T28S) the ability to form channels and either eliminate (T28G) or slow (T28V, T28S) the kinetics of productive RNA release (17). These observations have led to the hypothesis that the viral polypeptides insert into host cell membranes during infection and rearrange to form channels that permit the viral genome to pass through the membrane, thereby gaining access to the cytoplasm (7, 17, 49, 50).Speculation about the sites of externalization of the viral peptides and of the viral genome began soon after the structures of mature rhinovirus and poliovirus were determined crystallographically 25 years ago (31, 44). In both structures there is a solvent-filled channel running along each 5-fold axis. This channel is closed off at the outer surface of the capsid by polypeptide loops and on the inner surface by a plug that is formed by five intertwined copies of the amino terminus of VP3, forming a parallel beta tube (31, 44). In poliovirus this tube is flanked on its inner surface by five copies of a three-stranded beta sheet in which the outermost two strands come from a beta hairpin at the amino terminus of VP4 and the innermost strand comes from residues at the extreme amino terminus of VP1 (20). The presence of this channel, together with its proximity to peptide segments that were known to be externalized upon receptor attachment, and analogies with other viruses led to a model in which both the peptides and the viral RNA are externalized via the channel at the 5-fold axis (25, 45). At that time, an alternative model for the egress of polypeptides was proposed, based on an analogy with the externalization of the amino-terminal extensions of capsid proteins in expanded states of the topologically similar T=3 plant viruses (26, 32, 43, 47) and on genetic and biochemical studies of mutations that affect cell entry and capsid stability in poliovirus (14, 39, 54). In the latter model, the peptides were proposed to exit from the base of the canyon and then proceed along the outer surface toward the 5-fold peak (43, 47). Both models suggested that five copies of each of the externalized peptides would interact in some way to form a pore in the membrane that was contiguous with one of the 5-fold channels, thus providing a way for RNA to be released from the virion at a 5-fold axis of symmetry. No data yet exist to specify what specific structural roles VP4 and the amino terminus of VP1 might play in forming pores and serving as membrane anchors. However, both the electrophysiology data (cited above) and the greater sequence conservation of VP4 suggest that its role in pore formation may be the more central (17, 49, 50).To further elucidate various steps along the infection pathway, cryo-electron microscopy (cryo-EM) reconstructions have been determined for a number of cell entry intermediates of poliovirus and rhinoviruses, and their resolutions have been improved over time (2, 3, 7, 28, 38). Structures of the complexes of polioviruses and major-group rhinoviruses with the ectodomains of their respective receptors have confirmed earlier models that suggested that the canyon is the receptor-binding site and have begun to suggest how receptor binding might lead to receptor-induced conformational rearrangements (3, 56). Cryo-EM and cryo-electron tomography structures (cryo-ET) of a poliovirus-receptor-membrane complex (using a novel receptor-decorated liposome model [51]) confirmed that initial receptor binding brings the surface of the 5-fold mesa into close proximity with the membrane and appears to produce an outward distortion of the outer leaflet of the membrane in its area of closest approach to the virus particle (4, 8).Structures have also been determined for the soluble 135S and 80S particles of poliovirus, formed by heating the virus at 50°C (135S) or 56°C (80S) in hypotonic buffers, and for the 80S particles of rhinovirus 14 and 16, formed by exposing virus to acidic pH. All of the biological and immunological evidence that is currently available indicates that the particles prepared in vitro and used for structural studies are indistinguishable from the particles that are released from the cell surface during infection (6, 53). These structures have allowed the models for peptide release and genome release to be extended and refined (7, 38) and indeed have confirmed that VP1 exits from the particle surface at the base of the canyon and climbs up the side of the 5-fold mesa. However, contrary to the assumptions of the earlier models, the 10-Å structures of the poliovirus 135S and 80S particles show that the amino end of the amino-terminal extension of VP1 does not remain associated with the mesa. Instead, it forms an alpha-helical bridge that stretches across the canyon and binds to the large EF loop of VP2, a surface projection that appears as a 3-fold propeller blade (7, 38).Until recently, the mechanism of RNA release (during the 135S-to-80S transition) has been largely a matter of conjecture. We can infer that the RNA must exit via a single site on the virion surface, to avoid entanglement with the capsid (particularly as entanglement has never been observed in electron micrographs), though the mechanism by which a single site is selected (from among 60 equivalents) is unknown. All models presented to date have assumed that the RNA is released from the channel at the 5-fold axes (2, 3, 7, 8, 25, 27, 28, 30, 42, 45). However, in the icosahedrally constrained 10-Å structures of both the poliovirus 135S and 80S particles (7, 38), the apparent intactness and stability of the 5-fold mesa argues against the 5-fold axis being the site of RNA egress, given that the diameter of the opening, as seen in those structures, would be insufficient to accommodate RNA, even if the “plug” formed by the intertwined amino termini of VP3 was displaced. Moreover, both structures revealed significant thinning between 2-fold-related pentamers in the vicinity of the 2-fold axes. Most convincingly, large holes (easily sufficient to accommodate RNA) were seen at and near the 2-fold axes in the atomic model of the late-80S structure. This coincided with an open hole in the reconstruction, when viewed at a contour level that left most of the remainder of the capsid intact. This evidence was suggestive, but not definitive, as a number of other openings were present, particularly in the interfaces between protomers. Furthermore, the behavior of the capsid structure in the immediate vicinity of the unique site of RNA exit is likely to be different from what we see in the icosahedral average, which is dominated by the remainder of the capsid.In the course of solving icosahedrally symmetric cryo-EM structures for the poliovirus end-state 80S empty capsid particle (7, 38), we were surprised to find that RNA was frequently visible near the capsid and that in a subset of about 5% of the sampled virions, RNA was seen on both the inside and outside of the capsid, apparently caught in the act of exiting. This was an exciting development, as images of viral RNA release had never previously been reported. We were able to improve the resolution to ∼10 Å by classifying the projected images into two groups: an early 80Se particle that was more prevalent in the population after a shorter heating time and a late 80Sl particle that was seen more often when the heating time was increased. The amount of RNA density remaining in the interior appears to be continuously variable in both classes, suggesting that release is gradual. Of the 5% subset of particles clearly caught in the act, almost all belonged to the 80Se class. Our interpretation was that the 80Se class may represent particles in which exiting RNA is still engaged with the capsid machinery and traversing the capsid, while the 80Sl class (in which much of the capsid resembles the 135S form more closely in structure) represents particles with the RNA disengaged, possibly after nuclease cleavage. More than two structural classes may be present, but at the current resolution, we could not distinguish them.The present report addresses the question of what we can learn about the details of RNA release from an asymmetric cryo-EM reconstruction, based on the 540-particle caught-in-the-act subset, and independently from cryo-electron tomographic reconstructions of a similarly prepared sample. In each projected particle image or subtomogram, preliminary orientation parameters are first determined from an icosahedrally symmetric calculation, and in a second stage, the symmetry is broken by choosing 1 of the 60 symmetry-equivalent orientations. Both methods have yielded similar information, at about 50-Å resolution, concerning the footprint of the RNA on the virion surface, which demonstrates that RNA is released from an asymmetric site at the base of the canyon near a particle 2-fold axis and not at the channel at the 5-fold axes, as suggested by previous models. Additionally, the demonstrated success of the methodology provides us with a blueprint for resolving the molecular details of the RNA-capsid interaction in future experiments.  相似文献   

13.
A crystal structure has been obtained for a 26-nucleotide RNA that contains the loop E sequence from Chromatium minutissimum. Rather than having a loop E-like conformation, it consists of an A-form helix that splits into two separate strands following a sheared A-G base pair. The backbone of the strand containing the G of the A-G pair makes a turn of almost 180 degrees in the space of two nucleotides, and then interacts with the minor groove of the helix from which it originates. Similar structures, which we call hook-turns, occur in 16S and 23S rRNAs. They are found at places where the two strands of a helix separate at an A/G juxtaposition to interact with other sequences.  相似文献   

14.
BACKGROUND: Spo0F and Spo0B specifically exchange a phosphoryl group in a central step of the phosphorelay signal transduction system that controls sporulation in Bacilli. Spo0F belongs to the superfamily of response regulator proteins and is one of 34 such proteins in Bacillus subtilis. Spo0B is structurally similar to the phosphohistidine domain of histidine kinases, such as EnvZ, and exchanges a phosphoryl group between His30 and Asp54 on Spo0F. Information at the molecular level on the interaction between response regulators and phosphohistidine domains is necessary to develop a rationale for how phospho-signaling fidelity is maintained in two-component systems. RESULTS: Structural analysis of a co-crystal of the Spo0F response regulator interacting with the Spo0B phosphotransferase of the phosphorelay signal transduction system of B. subtilis was carried out using X-ray crystallographic techniques. The association of the two molecules brings the catalytic residues from both proteins into precise alignment for phosphoryltransfer. Upon complex formation, the Spo0B conformation remains unchanged. Spo0F also retains the overall conformation; however, two loops around the active site show significant deviations. CONCLUSIONS: The Spo0F-Spo0B interaction appears to be a prototype for response regulator-histidine kinase interactions. The primary contact surface between these two proteins is formed by hydrophobic regions in both proteins. The Spo0F residues making up the hydrophobic patch are very similar in all response regulators suggesting that the binding is initiated through the same residues in all interacting response regulator-kinase pairs. The bulk of the interactions outside this patch are through nonconserved residues. Recognition specificity is proposed to arise from interactions of the nonconserved residues, especially the hypervariable residues of the beta4-alpha4 loop.  相似文献   

15.

Background  

Difficult problems in structural bioinformatics are often studied in simple exact models to gain insights and to derive general principles. Protein folding, for example, has long been studied in the lattice model. Recently, researchers have also begun to apply the lattice model to the study of RNA folding.  相似文献   

16.
A new series of complexes of cobalt(II) fluoride, nickel(II) fluoride, copper(II) fluoride and zinc(II) fluoride with imidazole were synthesized and characterized by elemental analysis, molar conductance, magnetic moments, IR and electronic absorption measurements. Based on elemental and spectral data, the complexes were found to be of [M(im)6]F2 · XH2O type, where M is Co(II), Ni(II), Cu(II) and Zn(II) and X 4-5. The magnetic moments and spectral data suggested that all the complexes possessed an octahedral geometry. The crystal structure of the nickel complex, [Ni(im)6]F2 · 5H2O, is also reported in which nickel atom is surrounded by six nitrogen atoms of imidazole. Strong intra- and inter-molecular hydrogen bonding exists between fluoride ions (uncoordinated), nitrogen of imidazole and the -OH of water molecules.  相似文献   

17.
C L Liao  M M Lai 《Journal of virology》1992,66(10):6117-6124
Mouse hepatitis virus (MHV), a coronavirus, has been shown to undergo a high frequency of RNA recombination both in tissue culture and in animal infection. So far, RNA recombination has been demonstrated only between genomic RNAs of two coinfecting viruses. To understand the mechanism of RNA recombination and to further explore the potential of RNA recombination, we studied whether recombination could occur between a replicating MHV RNA and transfected RNA fragments. We first used RNA fragments which represented the 5' end of genomic-sense sequences of MHV RNA for transfection. By using polymerase chain reaction amplification with two specific primers, we were able to detect recombinant RNAs which incorporated the transfected fragment into the 5' end of the viral RNA in the infected cells. Surprisingly, even the anti-genomic-sense RNA fragments complementary to the 5' end of MHV genomic RNA could also recombine with the MHV genomic RNAs. This observation suggests that RNA recombination can occur during both positive- and negative-strand RNA synthesis. Furthermore, the recombinant RNAs could be detected in the virion released from the infected cells even after several passages of virus in tissue culture cells, indicating that these recombinant RNAs represented functional virion RNAs. The crossover sites of these recombinants were detected throughout the transfected RNA fragments. However, when an RNA fragment with a nine-nucleotide (CUUUAUAAA) deletion immediately downstream of a pentanucleotide (UCUAA) repeat sequence in the leader RNA was transfected into MHV-infected cells, most of the recombinants between this RNA and the MHV genome contained crossover sites near this pentanucleotide repeat sequence. In contrast, when exogenous RNAs with the intact nine-nucleotide sequence were used in similar experiments, the crossover sites of recombinants in viral genomic RNA could be detected at more-downstream sites. This study demonstrated that recombination can occur between replicating MHV RNAs and RNA fragments which do not replicate, suggesting the potential of RNA recombination for genetic engineering.  相似文献   

18.
A protein crystal lattice consists of surface contact regions, where the interactions of specific groups play a key role in stabilizing the regular arrangement of the protein molecules. In an attempt to control protein incorporation in a crystal lattice, a leucine zipper-like hydrophobic interface (comprising four leucine residues) was introduced into a helical region (helix 2) of the human pancreatic ribonuclease 1 (RNase 1) that was predicted to form a suitable crystallization interface. Although crystallization of wild-type RNase 1 has not yet been reported, the RNase 1 mutant having four leucines (4L-RNase 1) was successfully crystallized under several different conditions. The crystal structures were subsequently determined by X-ray crystallography by molecular replacement using the structure of bovine RNase A. The overall structure of 4L-RNase 1 is quite similar to that of the bovine RNase A, and the introduced leucine residues formed the designed crystal interface. To characterize the role of the introduced leucine residues in crystallization of RNase 1 further, the number of leucines was reduced to three or two (3L- and 2L-RNase 1, respectively). Both mutants crystallized and a similar hydrophobic interface as in 4L-RNase 1 was observed. A related approach to engineer crystal contacts at helix 3 of RNase 1 (N4L-RNase 1) was also evaluated. N4L-RNase 1 also successfully crystallized and formed the expected hydrophobic packing interface. These results suggest that appropriate introduction of a leucine zipper-like hydrophobic interface can promote intermolecular symmetry for more efficient protein crystallization in crystal lattice engineering efforts.  相似文献   

19.
Sequence and symmetry in ribosome binding sites of bacteriophage f1 RNA   总被引:20,自引:0,他引:20  
RNA was synthesized in vitro from α-32P-labeled ribonucleoside triphosphates with Escherichia coli RNA polymerase from covalently closed, circular, double-stranded DNA isolated from cells infected with bacteriophage f1. This RNA, which serves as an efficient message in vitro, was bound to ribosomes and the initiation complexes were digested with bovine pancreatic ribonuclease. Ribosome-protected fragments were isolated, purified and separated by two-dimensional analysis using electrophoresis and homochromatography. Sequence analysis, taking advantage of the ability to determine nearest neighbors, was done by conventional techniques.The sequences of the ribosome-protected fragments were found to fall into three classes. One sequence corresponds to the amino-terminal region of the protein product of f1 gene V, a DNA binding protein. It is proposed that a second sequence may correspond to the amino-terminal region of a precursor to the major coat protein. No assignment has yet been made for the third sequence.Comparisons are made between these three sequences and others that are available, both in terms of sequence features that have been pointed out earlier, and in terms of certain considerations of symmetry and syntax2 prominent in these binding site sequences that have not been discussed before.  相似文献   

20.
Muconolactone isomerase from Pseudomonas putida crystallizes from sodium sulfate solution in space group P2(1) (a = 65.84 A, b = 105.70 A, c = 77.20 A, beta = 90.5 degrees) with ten 11,000 Mr subunits per asymmetric unit. The 7 A resolution crystal structure was solved by single isomorphous replacement followed by 10-fold symmetry averaging. The decameric enzyme has an uncommon non-crystallographic 5-fold symmetry axis and a large cavity in its center.  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