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 83 毫秒
1.
Da LT  Quan JM  Wu YD 《Proteins》2011,79(6):1810-1819
A recently discovered small-molecule inhibitor, BMS-488043 (BMS-488), for the invasion of Human immunodeficiency virus Type 1 (HIV-1), shows a high activity against the entry of diversified HIV-1. Docking and molecular dynamic studies have been carried out to understand the binding mode of BMS-488 to gp120 as well as the effect of the small molecule on the conformational change of gp120 induced by CD4 binding. The results indicate that BMS-488 can accommodate in the CD4 binding pocket and interfere the CD4 binding in a noncompetitive mode. The piperazine group of BMS-488 prevents the bridging sheet formation of gp120 induced by the CD4 binding mainly through blocking the rotation of the Trp112 located on the α1 helix of gp120. The bridging sheet formation cannot be blocked for the W112A mutant of gp120 due to the reduced steric hindrance, in agreement with its significant resistance to the BMS inhibitor. The aza-indole ring is likely to interfere the exposure of gp41 by stacking within the β3-β5 and LB loops to disrupt the close packing of Pro212-His66-Phe210. The mode of action of BMS-488 also accommodates many mutagenesis results related to BMS-488 activity.  相似文献   

2.
Inappropriate activation of the Hh signaling pathway has been implicated in the development of several types of cancers including prostate, lung, pancreas, breast, brain and skin. Present study identified the binding affinities of eight established inhibitors viz., Cyclopamine, Saridegib, Itraconazole, LDE-225, TAK-441, BMS-833923 (XL139), PF-04449913 and Vismodegib targeting SMO receptor - a candidate protein involved in hedgehog pathway and sought to identify the best amongst the established inhibitors through by molecular docking. Exelxis® BMS 833923 (XL 139) demonstrated superior binding affinity aided by MolDock scoring docking algorithm. Further BMS 833923 (XL 139) was evaluated for pharmacophoric features which revealed appreciable ligand receptor interactions.  相似文献   

3.
HIV‐1 gp120 undergoes multiple conformational changes both before and after binding to the host CD4 receptor. BMS‐626529 is an attachment inhibitor (AI) in clinical development (administered as prodrug BMS‐663068) that binds to HIV‐1 gp120. To investigate the mechanism of action of this new class of antiretroviral compounds, we constructed homology models of unliganded HIV‐1 gp120 (UNLIG), a pre‐CD4 binding‐intermediate conformation (pCD4), a CD4 bound‐intermediate conformation (bCD4), and a CD4/co‐receptor‐bound gp120 (LIG) from a series of partial structures. We also describe a simple pathway illustrating the transition between these four states. Guided by the positions of BMS‐626529 resistance substitutions and structure–activity relationship data for the AI series, putative binding sites for BMS‐626529 were identified, supported by biochemical and biophysical data. BMS‐626529 was docked into the UNLIG model and molecular dynamics simulations were used to demonstrate the thermodynamic stability of the different gp120 UNLIG/BMS‐626529 models. We propose that BMS‐626529 binds to the UNLIG conformation of gp120 within the structurally conserved outer domain, under the antiparallel β20–β21 sheet, and adjacent to the CD4 binding loop. Through this binding mode, BMS‐626529 can inhibit both CD4‐induced and CD4‐independent formation of the “open state” four‐stranded gp120 bridging sheet, and the subsequent formation and exposure of the chemokine co‐receptor binding site. This unique mechanism of action prevents the initial interaction of HIV‐1 with the host CD4+ T cell, and subsequent HIV‐1 binding and entry. Our findings clarify the novel mechanism of BMS‐626529, supporting its ongoing clinical development. Proteins 2015; 83:331–350. © 2014 Wiley Periodicals, Inc.  相似文献   

4.
Crosslinking of monomeric IgG2 molecules bound to the Fc gamma receptors on the cell surface of guinea pig macrophages generated the triggering signal for the superoxide-generating system. A binding experiment indicated that macrophages have saturable binding sites for monomeric IgG2. Scatchard analysis of the binding data showed that macrophages have an average of 4 X 10(5) binding sites per cell and the association constant for the binding was 4.2 X 10(6) M-1. Binding of monomeric IgG2 to macrophages could be detected by subsequent reaction with the 125I-labeled F(ab')2 fragment of rabbit antibody specific for guinea pig Fab. Although binding of IgG2 monomer to Fc receptor did not stimulate superoxide release, further addition of the F(ab')2 fragment of anti-guinea pig Fab antibody did induce generation and release of superoxide, and the amount released was dependent on the dose of cell-bound IgG2. When macrophages were bound with a constant dose of IgG2 monomer in the first step, the superoxide release triggered by the addition of the F(ab')2 of anti-guinea pig Fab was dependent on the dose of the F(ab')2 fragment added. These results show that crosslinking of Fc receptors triggers the superoxide generation.  相似文献   

5.
6.
A molecular dynamics simulation (1.1 ns) at 300 K, of fully hydrated Ile21Cys, Glu25Cys plastocyanin mutant has been performed to investigate the structural, dynamical and functional effects of a disulfide bridge insertion at the surface of the protein. A detailed analysis of the root mean square fluctuations, H-bonding pattern and dynamical cross-correlation map has been performed. An essential dynamics method has also been applied as complementary analysis to identify concerted motions (essential modes), that could be relevant to the electron transfer function. The results have been compared with those previously obtained for wild-type plastocyanin and have revealed that the mutant shows a different pattern of H-bonds, with several interactions lost and a higher flexibility, especially around the electron transfer copper site. The analysis of dynamical cross-correlation map and of essential modes, has shown that the mutant performs different functional concerted motions, which might be related to the binding recognition with its electron transfer partners in comparison with the wild-type protein.  相似文献   

7.
Isolated influenza A virus nucleoprotein exists in an equilibrium between monomers and trimers. Samples containing only monomers or only trimers can be stabilized by respectively low and high salt. The trimers bind RNA with high affinity but remain trimmers, whereas the monomers polymerise onto RNA forming nucleoprotein-RNA complexes. When wild type (wt) nucleoprotein is crystallized, it forms trimers, whether one starts with monomers or trimers. We therefore crystallized the obligate monomeric R416A mutant nucleoprotein and observed how the domain exchange loop that leads over to a neighbouring protomer in the trimer structure interacts with equivalent sites on the mutant monomer surface, avoiding polymerisation. The C-terminus of the monomer is bound to the side of the RNA binding surface, lowering its positive charge. Biophysical characterization of the mutant and wild type monomeric proteins gives the same results, suggesting that the exchange domain is folded in the same way for the wild type protein. In a search for how monomeric wt nucleoprotein may be stabilized in the infected cell we determined the phosphorylation sites on nucleoprotein isolated from virus particles. We found that serine 165 was phosphorylated and conserved in all influenza A and B viruses. The S165D mutant that mimics phosphorylation is monomeric and displays a lowered affinity for RNA compared with wt monomeric NP. This suggests that phosphorylation may regulate the polymerisation state and RNA binding of nucleoprotein in the infected cell. The monomer structure could be used for finding new anti influenza drugs because compounds that stabilize the monomer may slow down viral infection.  相似文献   

8.
9.
The ATPase activity of the molecular chaperone Hsp90 is essential for its function in the assembly of client proteins. To understand the mechanism of human Hsp90, we have carried out a detailed kinetic analysis of ATP binding, hydrolysis and product release. ATP binds rapidly in a two-step process involving the formation of a diffusion-collision complex followed by a conformational change. The rate-determining step was shown to be ATP hydrolysis and not subsequent ADP dissociation. There was no evidence from any of the biophysical measurements for cooperativity in either nucleotide binding or hydrolysis for the dimeric protein. A monomeric fragment, lacking the C-terminal dimerisation domain, showed no dependence on protein concentration and, therefore, subunit association for activity. The thermodynamic linkage between client protein binding and nucleotide affinity revealed ATP bound Hsp90 has a higher affinity for client proteins than the ADP bound form. The kinetics are consistent with independent Michaelis-Menten catalysis in each subunit of the Hsp90 dimer. We propose that Hsp90 functions in an open-ring configuration for client protein activation.  相似文献   

10.
KpnI DNA-(N6-adenine) methyltransferase (M.KpnI) recognises the sequence 5'-GGTACC-3' and transfers the methyl group from S-adenosyl-L-methionine (AdoMet) to the N6 position of the adenine residue in each strand. Earlier studies have shown that M.KpnI exists as a dimer in solution, unlike most other MTases. To address the importance of dimerisation for enzyme function, a three-dimensional model of M.KpnI was obtained based on protein fold-recognition analysis, using the crystal structures of M.RsrI and M.MboIIA as templates. Residues I146, I161 and Y167, the side chains of which are present in the putative dimerisation interface in the model, were targeted for site-directed mutagenesis. Methylation and in vitro restriction assays showed that the mutant MTases are catalytically inactive. Mutation at the I146 position resulted in complete disruption of the dimer. The replacement of I146 led to drastically reduced DNA and cofactor binding. Substitution of I161 resulted in weakening of the interaction between monomers, leading to both monomeric and dimeric species. Steady-state fluorescence measurements showed that the wild-type KpnI MTase induces structural distortion in bound DNA, while the mutant MTases do not. The results establish that monomeric MTase is catalytically inactive and that dimerisation is an essential event for M.KpnI to catalyse the methyl transfer reaction.  相似文献   

11.
We investigated structural requirements for dimerisation and ligand binding of insulin/IGF receptors. Soluble receptor fragments consisting of N-terminal domains (L1/CYS/L2, L1/CYS/L2/F0) or fibronectin domains (F0/F1/F2, F1/F2) were expressed in CHO cells. Fragments containing F0 or F1 domains were secreted as disulphide-linked dimers, and those consisting of L1/CYS/L2 domains as monomers. None of these proteins bound ligand. However, when a peptide of 16 amino acids from the alpha-subunit C-terminus was fused to the C-terminus of L1/CYS/L2, the monomeric insulin and IGF receptor constructs bound their respective ligands with affinity only 10-fold lower than native receptors.  相似文献   

12.
Molecular dynamics (MD) simulations of the estrogen receptor DNA-binding domain (ERDBD) as a dimer in complex with its DNA response element (ERE) show a significant difference in both structure and dynamics, compared to a MD simulation of monomeric ERDBD bound to its half-site response element (EREH). The C-terminal zinc binding domain (ZnII), including a region (helix II) which is in a helical conformation in ERE-(ERDBD)2, is considerably more flexible in EREH-ERDBD than in the dimeric complex. In EREH-ERDBD, all helical hydrogen bonds in helix II are broken and the entire ZnII region is detached from a hydrogen bonding network that in ERE-(ERDBD)2 connects to other parts of the protein as well as to the DNA. The regions that become flexible in EREH-ERDBD are identical to the regions where the NMR solution structure of free ERDBD is poorly ordered. This strongly suggests that dimerisation of ERDBD is required for ordering of the ZnII region and that monomeric binding to DNA is not sufficient for the ordering. This contrasts to the glucocorticoid receptor DNA-binding domain (GRDBD) which has essentially the same mobility (uniform and limited), regardless of whether it is free as a monomer in solution, bound as a monomer to its half-site response element or in a dimeric complex with the full response element. The hydrogen bonding network that connects ZnII with other parts of the protein and to DNA is almost identical in ERDBD and GRDBD. However, in GRDBD there is also a serine (in the N-terminal zinc coordinating region) with a central role in this network, connecting to the ZnII region. This serine is replaced by a glycine in ERDBD and we suggest that this substitution is sufficient for destabilisation of the network, thus leading to a more flexible ZnII region, which becomes ordered first upon forming a complex with another ERDBD and DNA. Received: 6 March 1998 / Revised version: 22 June 1998 / Accepted: 2 September 1998  相似文献   

13.
Plasmodium falciparum triosephosphate isomerase (PfTIM) is known to be functional only as a homodimer. Although many studies have shown that the interface Cys13 plays a major role in the stability of the dimer, a few reports have demonstrated that structurally conserved Tyr74 may be essential for the stability of PfTIM dimer. To understand the role of Tyr74, we have performed molecular dynamics (MD) simulations of monomeric and dimeric PfTIM mutated to glycine and cysteine at position 74. Simulations of the monomer revealed that mutant Tyr74Gly does not produce changes in folding and stability of the monomer. Interestingly, comparison of the flexibility of Tyr74 in the monomer and dimer revealed that this residue possesses an intrinsic restricted mobility, indicating that Tyr74 is an anchor residue required for homodimerization. Tyr74 also appears to play an important role in binding by facilitating the disorder-to-order transitions of loops 1 and 3, which allows Cys13 to form favorable interactions with loop 3 and Lys12 to be locked in a favorable position for catalysis. High-temperature MD simulations of the wild-type and Tyr74Gly PfTIM dimers showed that the aromatic moiety of Tyr74 is necessary to preserve the geometry and native contacts between loops 1 and 3 at the interface of the dimer. Disulfide cross-linking between mutant Tyr74Cys and Cys13 further revealed that Tyr74 stabilizes the geometry of loop 1 (which contains the catalytic residue Lys12) and the interactions between loops 1 and 3 via aromatic-aromatic interactions with residues Phe69, Tyr101, and Phe102. Principal component analysis showed that Tyr74 is also necessary to preserve the collective motions in the dimer that contribute to the catalytic efficiency of PfTIM dimer. We conclude that Tyr74 not only plays a role in the stability of the dimer, but also participates in the dimerization process and collective motions via coupled disorder-to-order transitions of intrinsically disordered regions, necessary for efficiency in the catalytic function of PfTIM.  相似文献   

14.
Human peritoneal macrophages possess two populations of IgG Fc receptors   总被引:2,自引:0,他引:2  
To characterize the binding properties of the Fc receptors on human macrophages, the binding of radiolabeled human IgG1 to peritoneal macrophages was assessed. Cells were obtained at the time of diagnostic laparoscopy from women undergoing evaluation of infertility. Macrophages bound on the average more IgG1 monomer than monocytes but the avidity with which both types of cells bound IgG1 monomer was comparable. By contrast, macrophages bound much more IgG1 dimers than monocytes. Scatchard plots of the binding of dimer to monocytes were linear, but plots of binding to macrophages were markedly curvilinear. This curvilinearity was not an artifact of extensive ligand internalization or catabolism by cells, since 80% of binding was reversible and there was very little catabolism of ligand in the medium. Assuming that the observed curvilinearity was due to the presence of two independent subpopulations of receptors, an objective estimate for the number of receptors per cell and of the avidity with which each subpopulation bound IgG1 dimer was obtained using a previously described computer program (Scatfit). The analysis of the binding of dimer to macrophages from six donors suggested the presence of 42,000 +/- 33,000 high avidity receptors per cell which bind IgG1 dimer with a mean Ka of 2.7 X 10(9) M-1 and 218,000 +/- 127,000 low avidity receptors which bind the same ligand with a Ka of 1.1 X 10(7) M-1. ADCC of IgG antibody-coated sheep red blood cells mediated by macrophages was less readily inhibited by soluble IgG1 monomer than ADCC mediated by peripheral blood monocytes. This provides further evidence for the presence of low avidity receptors which bind monomeric IgG1 poorly and also suggests that these sites are functionally active in triggering antibody-dependent immune clearance.  相似文献   

15.
16.
R Koren  G G Hammes 《Biochemistry》1975,14(5):1021-1025
The interaction of NADH with s-malate dehydrogenase isolated from beef heart was studied in 20 mM potassium phosphate (pH 6.9)-1 mM EDTA, with forced dialysis, fluorescence, and temperature-jump techniques. Measurements of the change in fluorescence of NADH when it is titrated with enzyme indicate NADH bound to monomeric and dimeric enzyme have different fluorescence yields. These data and the results of direct binding studies can be explained in terms of a model in which the NADH binding sites on dimeric enzyme are equivalent or nearly equivalent, and NADH binding to monomeric enzyme occurs with an affinity very similar to that of the dimer. However, the fluorescence enhancement of NADH on binding to the enzyme is different for the monomer and for each of the two dimer sites.  相似文献   

17.
Staphylococcal α-hemolysin is expressed as a water-soluble monomeric protein and assembles on membranes to form a heptameric pore structure. The heptameric pore structure of α-hemolysin can be prepared from monomer in vitro only in the presence of deoxycholate detergent micelles, artificially constructed phospholipid bilayers, or erythrocytes. Here, we succeeded in preparing crystals of the heptameric form of α-hemolysin without any detergent but with 2-methyl-2,4-pentanediol (MPD), and determined its structure. The structure of the heptameric pore was similar to that reported previously. In the structure, two molecules of MPD were bound around Trp179, around which phospholipid head groups were bound in the heptameric pore structure reported previously. Size exclusion chromatography showed that α-hemolysin did not assemble spontaneously even when stored for 1 year. SDS-PAGE analysis revealed that, among the compounds in the crystallizing buffer, MPD could induce heptamer formation. The concentration of MPD that most efficiently induced oligomerization was between 10 and 30%. Based on these observations, we propose MPD as a reagent that can facilitate heptameric pore formation of α-hemolysin without membrane binding.  相似文献   

18.
Dystrophia myotonica protein kinase (DMPK) is a serine/threonine kinase composed of a kinase domain and a coiled‐coil domain involved in the multimerization. The crystal structure of the kinase domain of DMPK bound to the inhibitor bisindolylmaleimide VIII (BIM‐8) revealed a dimeric enzyme associated by a conserved dimerization domain. The affinity of dimerisation suggested that the kinase domain alone is insufficient for dimerisation in vivo and that the coiled‐coil domains are required for stable dimer formation. The kinase domain is in an active conformation, with a fully‐ordered and correctly positioned αC helix, and catalytic residues in a conformation competent for catalysis. The conserved hydrophobic motif at the C‐terminal extension of the kinase domain is bound to the N‐terminal lobe of the kinase domain, despite being unphosphorylated. Differences in the arrangement of the C‐terminal extension compared to the closely related Rho‐associated kinases include an altered PXXP motif, a different conformation and binding arrangement for the turn motif, and a different location for the conserved NFD motif. The BIM‐8 inhibitor occupies the ATP site and has similar binding mode as observed in PDK1.  相似文献   

19.

Background

The inhibitors blocking the interaction between programmed cell death protein 1(PD-1) and programmed death-ligand 1(PD-L1) can activate the immune response of T cell and eliminate cancer cells. The crystallographic studies have provided structural insights of the interactive interfaces between PD-L1 and its protein ligands. However, the hotspot residues on PD-L1 as well as structural and energetic basis for different protein ligands still need to be further investigated.

Methods

Molecular modeling methods including molecular dynamics simulation, per-residue free energy decomposition, virtual alanine scanning mutagenesis and residue-residue contact analysis were used to qualitatively and quantitatively analyze the interactions between PD-L1 and different protein ligands.

Results

The results of virtual alanine scanning mutagenesis suggest that Y56, Q66, M115, D122, Y123, R125 are the hotspot residues on PD-L1. The residue-residue contact analysis further shows that PD-1 interacts with PD-L1 mainly by F and G strands while monoclonal antibodies like avelumab and BMS-936559 mainly interact with PD-L1 by CDR2 and CDR3 loops of the heavy chain.

Conclusions

A structurally similar β-hairpin peptide with 13 or 14 residues was extracted from each protein ligand and these β-hairpin peptides were found tightly binding to the putative hotspot residues on PD-L1.

General significance

This study recognizes the hotspot residues on PD-L1 and uncovers the common structural and energetic basis of different protein ligands binding to PD-L1. These results will be valuable for the design of small molecule or peptide inhibitors targeting on PD-L1.  相似文献   

20.
Immunoglobulin binding domain B1 of streptococcal protein G (GB1), a small (56 residues), stable, single domain protein, is one of the most extensively used model systems in the area of protein folding and design. The recently determined NMR structure of a quadruple mutant (HS#124F26A, L5V/F30V/Y33F/A34F) revealed a domain-swapped dimer that dissociated into a partially folded, monomeric species at low micromolar protein concentrations. Here, we have characterized this monomeric, partially folded species by NMR and show that extensive conformational heterogeneity for a substantial portion of the polypeptide chain exists. Exchange between the conformers within the monomer ensemble on the microsecond to millisecond timescale renders the majority of backbone amide resonances broadened beyond detection. Despite these extensive temporal and spatial fluctuations, the overall architecture of the monomeric mutant protein resembles that of wild-type GB1 and not the monomer unit of the domain-swapped dimer.  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