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.
Foamy viruses belong to the genus Spumavirus of the family Retroviridae and have been isolated from many mammalian species. It was reported that simian foamy viruses (SFVs) have co-evolved with host species. In this study, we isolated four strains (WK1, WK2, AR1 and AR2) of SFV (named SFVjm) from Japanese macaques (Macaca fuscata) in main island Honshu of Japan. We constructed an infectious molecular clone of SFVjm strain WK1, termed pJM356. The virus derived from the clone replicated and induced syncytia in human (human embryonic kidney 293T cells), African green monkey (Vero cells) and mouse cell lines (Mus dunni tail fibroblast cells). Phylogenetic analysis also revealed that these four SFVjm strains formed two distinct SFVjm clusters. SFVjm strains WK1 and WK2 and SFV isolated from Taiwanese macaques (Macaca cyclopis) formed one cluster, whereas strains AR1 and AR2 formed the other cluster with SFV isolated from a rhesus macaque (Macaca mulatta).  相似文献   

2.
Efficient formation of specific intermolecular interactions is essential for self-assembly of biological structures. The foldon domain is an evolutionarily optimized trimerization module required for assembly of the large, trimeric structural protein fibritin from phage T4. Monomers consisting of the 27 amino acids comprising a single foldon domain subunit spontaneously form a natively folded trimer. During assembly of the foldon domain, a monomeric intermediate is formed on the submillisecond time scale, which provides the basis for two consecutive very fast association reactions. Mutation of an intermolecular salt bridge leads to a monomeric protein that resembles the kinetic intermediate in its spectroscopic properties. NMR spectroscopy revealed essentially native topology of the monomeric intermediate with defined hydrogen bonds and side-chain interactions but largely reduced stability compared to the native trimer. This structural preorganization leads to an asymmetric charge distribution on the surface that can direct rapid subunit recognition. The low stability of the intermediate allows a large free-energy gain upon trimerization, which serves as driving force for rapid assembly. These results indicate different free-energy landscapes for folding of small oligomeric proteins compared to monomeric proteins, which typically avoid the transient population of intermediates.  相似文献   

3.
Previously reported crystal structures of free and DNA-bound dimers of λ Cro differ strongly (about 4 Å backbone rmsd), suggesting both flexibility of the dimer interface and induced-fit protein structure changes caused by sequence-specific DNA binding. Here, we present two crystal structures, in space groups P3221 and C2 at 1.35 and 1.40 Å resolution, respectively, of a variant of λ Cro with three mutations in its recognition helix (Q27P/A29S/K32Q, or PSQ for short). One dimer structure (P3221; PSQ form 1) resembles the DNA-bound wild-type Cro dimer (1.0 Å backbone rmsd), while the other (C2; PSQ form 2) resembles neither unbound (3.6 Å) nor bound (2.4 Å) wild-type Cro. Both PSQ form 2 and unbound wild-type dimer crystals have a similar interdimer β-sheet interaction between the β1 strands at the edges of the dimer. In the former, an infinite, open β-structure along one crystal axis results, while in the latter, a closed tetrameric barrel is formed. Neither the DNA-bound wild-type structure nor PSQ form 1 contains these interdimer interactions. We propose that β-sheet superstructures resulting from crystal contact interactions distort Cro dimers from their preferred solution conformation, which actually resembles the DNA-bound structure. These results highlight the remarkable flexibility of λ Cro but also suggest that sequence-specific DNA binding may not induce large changes in the protein structure.  相似文献   

4.
No drug has been targeted specifically for HIV-2 (human immunodeficiency virus type 2) infection despite its increasing prevalence worldwide. The antiviral HIV-1 (human immunodeficiency virus type 1) protease (PR) inhibitor darunavir and the chemically related GRL98065 and GRL06579A were designed with the same chemical scaffold and different substituents at P2 and P2′ to optimize polar interactions for HIV-1 PR (PR1). These inhibitors are also effective antiviral agents for HIV-2-infected cells. Therefore, crystal structures of HIV-2 PR (PR2) complexes with the three inhibitors have been solved at 1.2-Å resolution to analyze the molecular basis for their antiviral potency. Unusually, the crystals were grown in imidazole and zinc acetate buffer, which formed interactions with the PR2 and the inhibitors. Overall, the structures were very similar to the corresponding inhibitor complexes of PR1 with an RMSD of 1.1 Å on main-chain atoms. Most hydrogen-bond and weaker C-H…O interactions with inhibitors were conserved in the PR2 and PR1 complexes, except for small changes in interactions with water or disordered side chains. Small differences were observed in the hydrophobic contacts for the darunavir complexes, in agreement with relative inhibition of the two PRs. These near-atomic-resolution crystal structures verify the inhibitor potency for PR1 and PR2 and will provide the basis for the development of antiviral inhibitors targeting PR2.  相似文献   

5.
The assignment of the 1H and 15Nnuclear magnetic resonance spectra of the Src-homology region 3 domain ofchicken brain -spectrin has been obtained. A set of solutionstructures has been determined from distance and dihedral angle restraints,which provide a reasonable representation of the protein structure insolution, as evaluated by a principal component analysis of the globalpairwise root-mean-square deviation (rmsd) in a large set of structuresconsisting of the refined and unrefined solution structures and the crystalstructure. The solution structure is well defined, with a lower degree ofconvergence between the structures in the loop regions than in the secondarystructure elements. The average pairwise rmsd between the 15 refinedsolution structures is 0.71 ± 0.13 Å for the backbone atoms and1.43 ± 0.14 Å for all heavy atoms. The solution structure isbasically the same as the crystal structure. The average rmsd between the 15refined solution structures and the crystal structure is 0.76 Å forthe backbone atoms and 1.45 ± 0.09 Å for all heavy atoms. Thereare, however, small differences probably caused by intermolecular contactsin the crystal structure.  相似文献   

6.
7.
8.
Foamy viruses (FV) are unusual retroviruses that differ in many aspects of their life cycle from the orthoretroviruses such as human immunodeficiency virus. Similar to Mason–Pfizer monkey virus (MPMV), FV assemble into capsids intracellularly. The capsids are then transported to a cellular membrane for acquisition of envelope (Env) glycoproteins and budding. However, unlike MPMV, budding of FV is dependent upon the presence of Env. Previous work suggested that FV Env proteins are localized to the endoplasmic reticulum (ER) where budding takes place. However, very little was known about the details of FV assembly. We have used immunofluorescence and electron microscopy to visualize the intracellular location of FV assembly and budding. We have found that, as in the case of MPMV, FV capsids assemble at a pericentriolar site in the cytoplasm. Surprisingly, FV Env is mostly absent from this site and, contrary to expectations, FV capsid structural protein (Gag) is absent from the ER. Gag and Env only co-localize at the trans -Golgi network, suggesting that Env–Gag interactions that are required for viral egress from the cell, occurs at this site. Finally, inhibitor studies suggest an important role of microtubule networks for foamy viral assembly and budding.  相似文献   

9.
10.
Pyridoxine 4-oxidase (PNOX) from Mesorhizobium loti is a monomeric glucose–methanol–choline (GMC) oxidoreductase family enzyme, catalyzes FAD-dependent oxidation of pyridoxine (PN) into pyridoxal, and is the first enzyme in pathway I for the degradation of PN. The tertiary structures of PNOX with a C-terminal His6-tag and PNOX–pyridoxamine (PM) complex were determined at 2.2 Å and at 2.1 Å resolutions, respectively. The overall structure consisted of FAD-binding and substrate-binding domains. In the active site, His460, His462, and Pro504 were located on the re-face of the isoalloxazine ring of FAD. PM binds to the active site through several hydrogen bonds. The side chains of His462 and His460 are located at 2.7 and 3.1 Å from the N4′ atom of PM. The activities of His460Ala and His462Ala mutant PNOXs were very low, and 460Ala/His462Ala double mutant PNOX exhibited no activity. His462 may act as a general base for the abstraction of a proton from the 4′-hydroxyl of PN. His460 may play a role in the binding and positioning of PN. The C4′ atom in PM is located at 3.2 Å, and the hydride ion from the C4′ atom may be transferred to the N5 atom of the isoalloxazine ring. The comparison of active site residues in GMC oxidoreductase shows that Pro504 in PNOX corresponds to Asn or His of the conserved His–Asn or His–His pair in other GMC oxidoreductases. The function of the novel proline residue was discussed.  相似文献   

11.
Structure-based drug design is underway to inhibit the S100B-p53 interaction as a strategy for treating malignant melanoma. X-ray crystallography was used here to characterize an interaction between Ca2+-S100B and TRTK-12, a target that binds to the p53-binding site on S100B. The structures of Ca2+-S100B (1.5-Å resolution) and S100B-Ca2+-TRTK-12 (2.0-Å resolution) determined here indicate that the S100B-Ca2+-TRTK-12 complex is dominated by an interaction between Trp7 of TRTK-12 and a hydrophobic binding pocket exposed on Ca2+-S100B involving residues in helices 2 and 3 and loop 2. As with an S100B-Ca2+-p53 peptide complex, TRTK-12 binding to Ca2+-S100B was found to increase the protein's Ca2+-binding affinity. One explanation for this effect was that peptide binding introduced a structural change that increased the number of Ca2+ ligands and/or improved the Ca2+ coordination geometry of S100B. This possibility was ruled out when the structures of S100B-Ca2+-TRTK-12 and S100B-Ca2+ were compared and calcium ion coordination by the protein was found to be nearly identical in both EF-hand calcium-binding domains (RMSD = 0.19). On the other hand, B-factors for residues in EF2 of Ca2+-S100B were found to be significantly lowered with TRTK-12 bound. This result is consistent with NMR 15N relaxation studies that showed that TRTK-12 binding eliminated dynamic properties observed in Ca2+-S100B. Such a loss of protein motion may also provide an explanation for how calcium-ion-binding affinity is increased upon binding a target. Lastly, it follows that any small-molecule inhibitor bound to Ca2+-S100B would also have to cause an increase in calcium-ion-binding affinity to be effective therapeutically inside a cell, so these data need to be considered in future drug design studies involving S100B.  相似文献   

12.
The presence of polyamines in the granulosis virus and its larval host, Plodia interpunctella, was examined using thin-layer chromatography of the dansylated compounds. The uninfected and infected larvae contain the three polyamines, putrescine, spermidine, and spermine. Although purified granulosis virus infecting this host does not contain any typical polyamines, an acidsoluble dansylated compound with properties indicative of a polyamine was observed.  相似文献   

13.
14.
A heme-acquisition system present in several Gram-negative bacteria requires the secretion of hemophores. These extracellular carrier proteins capture heme and deliver it to specific outer membrane receptors. The Serratia marcescens HasA hemophore is a monodomain protein that binds heme with a very high affinity. Its α/β structure, as that of its binding pocket, has no common features with other iron- or heme-binding proteins. Heme is held by two loops L1 and L2 and coordinated to iron by an unusual ligand pair, H32/Y75. Two independent regions of the hemophore β-sheet are involved in HasA-HasR receptor interaction. Here, we report the 3-D NMR structure of apoHasA and the backbone dynamics of both loaded and unloaded hemophore. While the overall structure of HasA is very similar in the apo and holo forms, the hemophore presents a transition from an open to a closed form upon ligand binding, through a large movement, of up to 30 Å, of loop L1 bearing H32. Comparison of loaded and unloaded HasA dynamics on different time scales reveals striking flexibility changes in the binding pocket. We propose a mechanism by which these structural and dynamic features provide the dual function of heme binding and release to the HasR receptor.  相似文献   

15.
Lactobacillus acidophilus NCFM is a probiotic bacterium known for its beneficial effects on human health. The importance of α-galactosidases (α-Gals) for growth of probiotic organisms on oligosaccharides of the raffinose family present in many foods is increasingly recognized. Here, the crystal structure of α-Gal from L. acidophilus NCFM (LaMel36A) of glycoside hydrolase (GH) family 36 (GH36) is determined by single-wavelength anomalous dispersion. In addition, a 1.58-Å-resolution crystallographic complex with α-d-galactose at substrate binding subsite − 1 was determined. LaMel36A has a large N-terminal twisted β-sandwich domain, connected by a long α-helix to the catalytic (β/α)8-barrel domain, and a C-terminal β-sheet domain. Four identical monomers form a tightly packed tetramer where three monomers contribute to the structural integrity of the active site in each monomer. Structural comparison of LaMel36A with the monomeric Thermotoga maritima α-Gal (TmGal36A) reveals that O2 of α-d-galactose in LaMel36A interacts with a backbone nitrogen in a glycine-rich loop of the catalytic domain, whereas the corresponding atom in TmGal36A is from a tryptophan side chain belonging to the N-terminal domain. Thus, two distinctly different structural motifs participate in substrate recognition. The tetrameric LaMel36A furthermore has a much deeper active site than the monomeric TmGal36A, which possibly modulates substrate specificity. Sequence analysis of GH36, inspired by the observed structural differences, results in four distinct subgroups having clearly different active-site sequence motifs. This novel subdivision incorporates functional and architectural features and may aid further biochemical and structural analyses within GH36.  相似文献   

16.
The X-ray structure of the C-terminal fragment, containing residues 449-946, of Escherichia coli glutamine synthetase adenylyl transferase (ATase) has been determined. ATase is part of the cascade that regulates the enzymatic activity of E. coli glutamine synthetase, a key component of the cell's machinery for the uptake of ammonia. It has two enzymatic activities, adenylyl removase (AR) and adenylyl transferase (AT), which are located in distinct catalytic domains that are separated by a regulatory (R) domain. We previously reported the three-dimensional structure of the AR domain (residues 1-440). The present structure contains both the R and AT domains. AR and AT share 24% sequence identity and also contain the β-polymerase motif that is characteristic of many nucleotidylyl transferase enzymes. The structures overlap with an rmsd of 2.4 Å when the superhelical R domain is omitted. A model for the complete ATase molecule is proposed, along with some refinements of domain boundaries. A rather more speculative model for the complex of ATase with glutamine synthetase and the nitrogen signal transduction protein PII is also presented.  相似文献   

17.
The structure and dynamic properties of an 80-residue fragment of Ste2p, the G-protein-coupled receptor for α-factor of Saccharomyces cerevisiae, was studied in LPPG micelles with the use of solution NMR spectroscopy. The fragment Ste2p(G31-T110) (TM1-TM2) consisted of 19 residues from the N-terminal domain, the first TM helix (TM1), the first cytoplasmic loop, the second TM helix (TM2), and seven residues from the first extracellular loop. Multidimensional NMR experiments on [15N], [15N, 13C], [15N, 13C, 2H]-labeled TM1-TM2 and on protein fragments selectively labeled at specific amino acid residues or protonated at selected methyl groups resulted in >95% assignment of backbone and side-chain nuclei. The NMR investigation revealed the secondary structure of specific residues of TM1-TM2. TALOS constraints and NOE connectivities were used to calculate a structure for TM1-TM2 that was highlighted by the presence of three α-helices encompassing residues 39-47, 49-72, and 80-103, with higher flexibility around the internal Arg58 site of TM1. RMSD values of individually superimposed helical segments 39-47, 49-72, and 80-103 were 0.25 ± 0.10 Å, 0.40 ± 0.13 Å, and 0.57 ± 0.19 Å, respectively. Several long-range interhelical connectivities supported the folding of TM1-TM2 into a tertiary structure typified by a crossed helix that splays apart toward the extracellular regions and contains considerable flexibility in the G56VRSG60 region. 15N-relaxation and hydrogen-deuterium exchange data support a stable fold for the TM parts of TM1-TM2, whereas the solvent-exposed segments are more flexible. The NMR structure is consistent with the results of biochemical experiments that identified the ligand-binding site within this region of the receptor.  相似文献   

18.
M-ficolin, which forms trimer-based multimers, is a pathogen-recognition protein in the innate immune system, and it binds to ligands through its fibrinogen-like (FBG) domain. As the first step toward the elucidation of the molecular basis for pathogen-recognition by the M-ficolin multimers, we assigned the backbone resonances of the monomeric mutant of the M-ficolin FBG domain, recombinantly expressed by Brevibacillus choshinensis. Like the wild-type trimeric FBG domain, the monomeric FBG domain also requires His251, His284 and His297 for the ligand-binding activity, as judged by mutational analyses using zonal affinity chromatography. The secondary structure predicted by the backbone resonance assignments is similar to that of the trimeric FBG domain in the crystal, indicating that the monomeric FBG domain is folded correctly to perform its function.  相似文献   

19.
The nuclear LSm2-8 (like Sm) complex and the cytoplasmic LSm1-7 complex play a central role in mRNA splicing and degradation, respectively. The LSm proteins are related to the spliceosomal Sm proteins that form a heteroheptameric ring around small nuclear RNA. The assembly process of the heptameric Sm complex is well established and involves several smaller Sm assembly intermediates. The assembly of the LSm complex, however, is less well studied. Here, we solved the 2.5 Å-resolution structure of the LSm assembly intermediate that contains LSm5, LSm6, and LSm7. The three monomers display the canonical Sm fold and arrange into a hexameric LSm657-657 ring. We show that the order of the LSm proteins within the ring is consistent with the order of the related SmE, SmF, and SmG proteins in the heptameric Sm ring. Nonetheless, differences in RNA binding pockets prevent the prediction of the nucleotide binding preferences of the LSm complexes. Using high-resolution NMR spectroscopy, we confirm that LSm5, LSm6, and LSm7 also assemble into a  60-kDa hexameric ring in solution. With a combination of pull-down and NMR experiments, we show that the LSm657 complex can incorporate LSm23 in order to assemble further towards native LSm rings. Interestingly, we find that the NMR spectra of the LSm57, LSm657-657, and LSm23-657 complexes differ significantly, suggesting that the angles between the LSm building blocks change depending on the ring size of the complex. In summary, our results identify LSm657 as a plastic and functional building block on the assembly route towards the LSm1-7 and LSm2-8 complexes.  相似文献   

20.
Retroviruses selectively package two copies of their RNA genomes via mechanisms that have yet to be fully deciphered. Recent studies with small fragments of the Moloney murine leukemia virus (MoMuLV) genome suggested that selection may be mediated by an RNA switch mechanism, in which conserved UCUG elements that are sequestered by base-pairing in the monomeric RNA become exposed upon dimerization to allow binding to the cognate nucleocapsid (NC) domains of the viral Gag proteins. Here we show that a large fragment of the MoMuLV 5′ untranslated region that contains all residues necessary for efficient RNA packaging (ΨWT; residues 147-623) also exhibits a dimerization-dependent affinity for NC, with the native dimer ([ΨWT]2) binding 12 ± 2 NC molecules with high affinity (Kd = 17 ± 7 nM) and with the monomer, stabilized by substitution of dimer-promoting loop residues with hairpin-stabilizing sequences (ΨM), binding 1-2 NC molecules. Identical dimer-inhibiting mutations in MoMuLV-based vectors significantly inhibit genome packaging in vivo (∼ 100-fold decrease), whereas a large deletion of nearly 200 nucleotides just upstream of the gag start codon has minimal effects. Our findings support the proposed RNA switch mechanism and further suggest that virus assembly may be initiated by a complex comprising as few as 12 Gag molecules bound to a dimeric packaging signal.  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