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 234 毫秒
1.
The conformation of oligomers of β‐amino acids of the general type Ac‐[β‐Xaa]n‐NHMe (β‐Xaa = β‐Ala, β‐Aib, and β‐Abu; n = 1–4) was systematically examined at different levels of ab initio molecular orbital theory (HF/6‐31G*, HF/3‐21G). The solvent influence was considered employing two quantum‐mechanical self‐consistent reaction field models. The results show a wide variety of possibilities for the formation of characteristic elements of secondary structure in β‐peptides. Most of them can be derived from the monomer units of blocked β‐peptides with n = 1. The stability and geometries of the β‐peptide structures are considerably influenced by the side‐chain positions, by the configurations at the Cα‐ and Cβ‐atoms of the β‐amino acid constituents, and especially by environmental effects. Structure peculiarities of β‐peptides, in particular those of various helix alternatives, are discussed in relation to typical elements of secondary structure in α‐peptides. © 1999 John Wiley & Sons, Inc. Biopoly 50: 167–184, 1999  相似文献   

2.
The villin headpiece helical subdomain (HP36) is one of the best known model systems for computational studies of fast‐folding all‐α miniproteins. HP21 is a peptide fragment—derived from HP36—comprising only the first and second helices of the full domain. Experimental studies showed that although HP21 is mostly unfolded in solution, it does maintain some persistent native‐like structure as indicated by the analysis of NMR‐derived chemical shifts. Here we compare the experimental data for HP21 with the results obtained from a 15‐μs long folding molecular dynamics simulation performed in explicit water and with full electrostatics. We find that the simulation is in good agreement with the experiment and faithfully reproduces the major experimental findings, namely that (a) HP21 is disordered in solution with <10% of the trajectory corresponding to transiently stable structures, (b) the most highly populated conformer is a native‐like structure with an RMSD from the corresponding portion of the HP36 crystal structure of <1 Å, (c) the simulation‐derived chemical shifts—over the whole length of the trajectory—are in reasonable agreement with the experiment giving reduced χ2 values of 1.6, 1.4, and 0.8 for the Δδ13Cα, Δδ13CO, and Δδ13Cβ secondary shifts, respectively (becoming 0.8, 0.7, and 0.3 when only the major peptide conformer is considered), and finally, (d) the secondary structure propensity scores are in very good agreement with the experiment and clearly indicate the higher stability of the first helix. We conclude that folding molecular dynamics simulations can be a useful tool for the structural characterization of even marginally stable peptides.  相似文献   

3.
Recent reports which show that several chemokines can act as direct microbicidal agents have drawn renewed attention to these chemotactic signalling proteins. Here we present a structure-function analysis of peptides derived from the human chemokines macrophage inflammatory protein-3α (MIP-3α/CCL20), interleukin-8 (IL-8), neutrophil activating protein-2 (NAP-2) and thrombocidin-1 (TC-1). These peptides encompass the C-terminal α-helices of these chemokines, which have been suggested to be important for the direct antimicrobial activities. Far-UV CD spectroscopy showed that the peptides are unstructured in aqueous solution and that a membrane mimetic solvent is required to induce a helical secondary structure. A co-solvent mixture was used to determine solution structures of the peptides by two-dimensional 1H-NMR spectroscopy. The highly cationic peptide, MIP-3α51-70, had the most pronounced antimicrobial activity and displayed an amphipathic structure. A shorter version of this peptide, MIP-3α59-70, remained antimicrobial but its structure and mechanism of action were unlike that of the former peptide. The NAP-2 and TC-1 proteins differ in their sequences only by the deletion of two C-terminal residues in TC-1, but intact TC-1 is a very potent antimicrobial while NAP-2 is inactive. The corresponding C-terminal peptides, NAP-250-70 and TC-150-68, had very limited and no bactericidal activity, respectively. This suggests that other regions of TC-1 contribute to its bactericidal activity. Altogether, this work provides a rational structural basis for the biological activities of these peptides and proteins and highlights the importance of experimental characterization of peptide fragments as distinct entities because their activities and structural properties may differ substantially from their parent proteins.  相似文献   

4.
A model of nine proteins including side-chain atoms have been built from the known Cα coordinates and amino acid sequences using a Monte Carlo Protein Building Annealing method. The Cartesian coordinates for the side-chain atoms were established with bond lengths and angles selected randomly from within previously determined ranges. A simulated annealing technique is used to generate some 300 structures with differing side-chain conformations. The atomic coordinates of the backbone atoms are fixed during the simulated annealing process. The coordinates of the side-chain atoms of 300 low energy conformations are averaged to obtain a mean structure that is minimized with the Cα atoms constrained to their position in the x-ray structure using the OPLS/AMBER force field with the GB/SA water model. The rms deviation of the main-chain atoms (without Cβ) compared with the corresponding crystal structures is in the range 0.20–0.64 Å. The rms deviation of the side-chain atoms is between 1.72 and 2.71 Å and for all atoms is between 1.19 and 1.99 Å. The method is insensitive to random errors in the Cα positions and the computational requirement is modest. © 1997 John Wiley & Sons, Inc.  相似文献   

5.
A series of terminally blocked peptides (to the pentamer level) from l ‐Ala and the cyclic Cα,α‐disubstituted Gly residue Afc and one Gly/Afc dipeptide have been synthesized by solution method and fully characterized. The molecular structure of the amino acid derivative Boc‐Afc‐OMe and the dipeptide Boc‐Afc‐Gly‐OMe were determined in the crystal state by X‐ray diffraction. In addition, the preferred conformation of all of the model peptides was assessed in deuterochloroform solution by FT‐IR absorption and 1H‐NMR. The experimental data favour the conclusion that the Afc residue tends to adopt either the fully‐extended (C5) or a folded/helical structure. In particular, the former conformation is highly populated in solution and is also that found in the crystal state in the two compounds investigated. A comparison with the structural propensities of the strictly related Cα,α‐disubstituted Gly residues Ac5c and Dϕg is made and the implications for the use of the Afc residue in conformationally constrained analogues of bioactive peptides are briefly examined. A spectroscopic (UV absorption, fluorescence, CD) characterization of this novel aromatic Cα,α‐disubstituted Gly residue is also reported. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
Conformational energy computations on a derivative and a homo-dipeptide of Cα,α-diethylglycine were performed. In both cases the N- and C-terminal groups are blocked as acetamido and methylamido moieties, respectively. It was found that the Cα,α-diethylglycine residues are conformationally restricted and that the minimum energy conformation corresponds to the fully extended C5 structure when the N? Cα? C′ bond angle is smaller than 108° (as experimentally observed). The results of the theoretical analysis are in agreement with the crystal-state structural propensity of the complete series of N-trifluoroacetylated homo-peptides of this Cα,α-dialkylated residue from monomer to pentamer, determined by x-ray diffraction and also described in this work. Interestingly, for the first time, a crystallographically planar peptide backbone was observed (in the protected tripeptide). A comparison with peptides of Cα,α-dimethylglycine, Cα-methyl, Cα-ethylglycine, and Cα,α-di-n-propylglycine indicates that the fully extended conformation becomes more stable than the helical structures when both amino acid side-chain Cβ atoms are substituted.  相似文献   

7.
β‐Peptides are analogs of natural α‐peptides and form a variety of remarkably stable structures. Having an additional carbon atom in the backbone of each residue, their folded conformation is not only influenced by the side‐chain sequence but also and foremost by their substitution pattern. The precise mechanism by which the side chains interact with the backbone is, however, hitherto not completely known. To unravel the various effects by which the side chains influence the backbone conformation, we quantify to which extent the dihedral angles of a β3‐substited peptide with an additional methyl group on the central Cα‐atom can be regarded as independent degrees of freedom and analyze the distributions of these dihedral angles. We also selectively capture the steric effect of substituents on the Cα‐ and Cβ‐atoms of the central residue by alchemically changing them into dummy atoms, which have no nonbonded interactions. We find that the folded state of the β3‐peptide is primarily stabilized by a steric exclusion of large parts of the unfolded state (entropic effect) and only subsequently by mutual dependence of the ψ‐dihedral angles (enthalpic effect). The folded state of β‐peptides is stabilized by a different mechanism than that of α‐peptides. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

8.
Mert Gur  Burak Erman 《Proteins》2012,80(12):2769-2779
Binding of two short peptides of sequences ASN‐ASP‐MET‐PHE‐ARG‐LEU and LEU‐LEU‐PHE‐MET‐GLN‐HIS and their bound complex structures is studied. Molecular dynamic simulations of the three structures around their respective minimum energy conformations are performed and a quasi‐harmonic analysis is performed over the trajectories generated. The fluctuation correlation matrix is constructed for all Cα‐atoms of the peptides for the full trajectory. The spring constant matrix between peptide Cα‐atoms is obtained from the correlation matrix. Statistical thermodynamics of fluctuations, the energies, entropies, and the free energies of binding are discussed in terms of the quasi‐harmonic model. Sites contributing to the stability of the system and presenting high affinity for binding are determined. Contribution of hydrophobic forces to binding is discussed. Quasi‐harmonic approximation identifies the essential subspace of motions, the important interactions, and binding sites, gives the energetic contribution of each individual interaction, and filters out noise observed in molecular dynamics owing to uncorrelated motions. Comparison of the molecular dynamics results with those of the quasi‐harmonic model shows the importance of entropy change, resulting from water molecules being liberated from the surfaces of the two peptides upon binding. Proteins 2012; © 2012 Wiley Periodicals, Inc.  相似文献   

9.
Min Zhang  Tuck C. Wong 《Biopolymers》1993,33(12):1901-1908
High-resolution proton spectra at 500 MHz of two tachykinin peptides, substance P methyl ester (SPOMe) and [Nle10]-neurokinin A (4–10), have been obtained in dimethylsulfoxide (DMSO), and for SPOMe, also in 2, 2, 2-trifluoroethanol (TFE)/water mixtures. Complete chemical shift assignments for these peptides were made based on two-dimensional (2D) nmr techniques, correlated spectroscopy and total COSY. J coupling measurement and nuclear Overhauser effect spectroscopy (NOESY) were then used to determine the conformation of these peptides in the various solvents. Based on the J coupling, NOE correlations, and temperature coefficients of the NH resonances, it is concluded that these two peptides exist in DMSO at room temperature as a mixture of conformers that are primarily extended. For SPOMe in TFE/water with high TFE content, however, helical structures are found to be present, and they become quite clear at temperatures between 270 and 280 K. The variation of the 13C chemical shifts of the Cα (the secondary shift) with TFE contents corroborates this conclusion. The NOE and Cα shifts show that the main helical region for SPOMe lies between 4P and 9G. The C-terminus segment L? M? NH2 is found to be quite flexible, which appears to be quite common for neurokinin-1 selective peptides. © 1993 John Wiley & Sons, Inc.  相似文献   

10.
Annexin A4 (Anx4) possesses four repeat domains with one Ca2+-binding site (CBS) in each domain. In this study, we resolved two crystal structures of the Na+-bound form at high resolution (1.58 and 1.35 Å). This is the first report that Anx4 binds the Na+ ion in CBSs. Electron density maps, valence screening, and atomic absorbance spectrometry confirmed that Anx4 bound the Na+ ion. One structure (1.58 Å) bound the Na+ ion in CBS I, whereas another structure (1.35 Å) bound the Na+ ion in CBS II and CBS III. We compared the two Na+-bound forms by superimposing their Cα traces. The Cα atoms of CBS III largely moved by coordination of the Na+ ion. In the Cα atoms of CBS I, however, little change resulted from Na+-coordination. Only the side chain of Glu71 was moved by Na+-coordination in CBS I. These results indicate that Anx4 also binds not only Ca2+ but also Na+ ion in the CBS.  相似文献   

11.
α‐Amino acid residues with a ?,ψ constrained conformation are known to significantly bias the peptide backbone 3D structure. An intriguing member of this class of compounds is (αMe)Aze, characterized by an Nα‐alkylated four‐membered ring and Cα‐methylation. We have already reported that (S)‐(αMe)Aze, when followed by (S)‐Ala in the homochiral dipeptide sequential motif ‐(S)‐(αMe)Aze‐(S)‐Ala‐, tends to generate the unprecedented γ‐bend ribbon conformation, as formation of a regular, fully intramolecularly H‐bonded γ‐helix is precluded, due to the occurrence of a tertiary amide bond every two residues. In this work, we have expanded this study to the preparation and 3D structural analysis of the heterochiral (S)‐Ala/(R)‐(αMe)Aze sequential peptides from dimer to hexamer. Our conformational results show that members of this series may fold in type‐II β‐turns or in γ‐turns depending on the experimental conditions.  相似文献   

12.
A fundamental problem in biochemistry and molecular biology is understanding the spatial structure of macromolecules and then analyzing their functions. In this study, the three-dimensional structure of a ribosome-inactivating protein luffin-α was predicted using a neural network method and molecular dynamics simulation. A feedforward neural network with the backpropagation learning algorithm were trained on model class of homologous proteins including trichosanthin andα-momorcharin. The distance constraints for the Cα atoms in the protein backbone were utilized to generate a folded crude conformation of luffin-α by model building and the steepest descent minimization approach. The crude conformation was refined by molecular dynamics techniques and a simulated annealing procedure. The interaction between luffin-α and its analogous substrate GAGA was also simulated to understand its action mechanism.  相似文献   

13.
Protein chemical shifts have long been used by NMR spectroscopists to assist with secondary structure assignment and to provide useful distance and torsion angle constraint data for structure determination. One of the most widely used methods for secondary structure identification is called the Chemical Shift Index (CSI). The CSI method uses a simple digital chemical shift filter to locate secondary structures along the protein chain using backbone 13C and 1H chemical shifts. While the CSI method is simple to use and easy to implement, it is only about 75–80 % accurate. Here we describe a significantly improved version of the CSI (2.0) that uses machine-learning techniques to combine all six backbone chemical shifts (13Cα, 13Cβ, 13C, 15N, 1HN, 1Hα) with sequence-derived features to perform far more accurate secondary structure identification. Our tests indicate that CSI 2.0 achieved an average identification accuracy (Q3) of 90.56 % for a training set of 181 proteins in a repeated tenfold cross-validation and 89.35 % for a test set of 59 proteins. This represents a significant improvement over other state-of-the-art chemical shift-based methods. In particular, the level of performance of CSI 2.0 is equal to that of standard methods, such as DSSP and STRIDE, used to identify secondary structures via 3D coordinate data. This suggests that CSI 2.0 could be used both in providing accurate NMR constraint data in the early stages of protein structure determination as well as in defining secondary structure locations in the final protein model(s). A CSI 2.0 web server (http://csi.wishartlab.com) is available for submitting the input queries for secondary structure identification.  相似文献   

14.
M. T. Cung  M. Marraud 《Biopolymers》1982,21(5):953-967
We use the nmr data concerning the CαH? CβH fragment in eight peptides with rigid side chains to parametrize a Karplus correlation between the vicinal proton Jαβ coupling constant and the dihedral angle θ. When considering molecules containing the fragment CαHα? CβHβHβ′, the three-dimensional structure of the model peptides does not need to be known with accurate precision, since each set of Jαβ and Jαβ′ coupling constants is then related to the coefficients of the Karplus equation. A good correlation is observed with the Karplus equation, which is in substantial agreement with the Jαβ coupling constants reported for rigid as well as rotating Cα? Cβ bonds in peptides.  相似文献   

15.
The solid‐state conformations of two αγ hybrid peptides Boc‐[Aib‐γ4(R)Ile]4‐OMe 1 and Boc‐[Aib‐γ4(R)Ile]5‐OMe 2 are described. Peptides 1 and 2 adopt C12‐helical conformations in crystals. The structure of octapeptide 1 is stabilized by six intramolecular 4 → 1 hydrogen bonds, forming 12 atom C12 motifs. The structure of peptide 2 reveals the formation of eight successive C12 hydrogen‐bonded turns. Average backbone dihedral angles for αγ C12 helices are peptide 1 , Aib; φ (°) = ?57.2 ± 0.8, ψ (°) = ?44.5 ± 4.7; γ4(R)Ile; φ (°) = ?127.3 ± 7.3, θ1 (°) = 58.5 ± 12.1, θ2 (°) = 67.6 ± 10.1, ψ (°) = ?126.2 ± 16.1; peptide 2 , Aib; φ (°) = ?58.8 ± 5.1, ψ (°) = ?40.3 ± 5.5; ψ4(R)Ile; φ (°) = ?123.9 ± 2.7, θ1 (°) = 53.3 θ 4.9, θ 2 (°) = 61.2 ± 1.6, ψ (°) = ?121.8 ± 5.1. The tendency of γ4‐substituted residues to adopt gauche–gauche conformations about the Cα–Cβ and Cβ–Cγ bonds facilitates helical folding. The αγ C12 helix is a backbone expanded analog of α peptide 310 helix. The hydrogen bond parameters for α peptide 310 and α‐helices are compared with those for αγ hybrid C12 helix. Copyright © 2016 European Peptide Society and John Wiley & Sons.  相似文献   

16.
The crystal-state preferred conformations of six Nα-blocked pentapeptide esters, each containing four helicogenic, achiral α-aminoisobutyric acid (Aib) residues followed by one chiral L -valine (L -Val) or Cα-methyl-L -valine [(αMe)Val] residue at the C-terminus, have been assessed by x-ray diffraction analysis. In all of the compounds the  (Aib)4 sequence is folded in a regular 310-helical conformation. In the four pentapeptides characterized by the L -(αMe)Val residue two conformationally distinct molecules occur in the asymmetric unit. Conversely, only one molecule is observed in the asymmetric unit of two pentapeptides with the C-terminal L -Val residue. In the L -Val based peptides the helical screw sense of the  (Aib)4 sequence is right-handed, whereas in the L  (αMe)Val analogues both right- and left-handed helical screw senses concomitantly occur in the two crystallographically independent molecules. © 1998 John Wiley & Sons, Inc. Biopoly 46: 433–443, 1998  相似文献   

17.
Background

A commonly recurring problem in structural protein studies, is the determination of all heavy atom positions from the knowledge of the central α-carbon coordinates.

Results

We employ advances in virtual reality to address the problem. The outcome is a 3D visualisation based technique where all the heavy backbone and side chain atoms are treated on equal footing, in terms of the Cα coordinates. Each heavy atom is visualised on the surfaces of a different two-sphere, that is centered at another heavy backbone and side chain atoms. In particular, the rotamers are visible as clusters, that display a clear and strong dependence on the underlying backbone secondary structure.

Conclusions

We demonstrate that there is a clear interdependence between rotameric states and secondary structure. Our method easily detects those atoms in a crystallographic protein structure which are either outliers or have been likely misplaced, possibly due to radiation damage. Our approach forms a basis for the development of a new generation, visualization based side chain construction, validation and refinement tools. The heavy atom positions are identified in a manner which accounts for the secondary structure environment, leading to improved accuracy.

  相似文献   

18.
The backbone of a protein is typically represented as either a C α-polyline, a three-dimensional (3D) polyline that passes through the C α atoms, or a tuple of ϕ,ψ pairs while its fold is usually assigned using the 3D topological arrangement of the secondary structure elements (SSEs). It is tricky to obtain the SSE composition for a protein from the C α-polyline representation while its 3D SSE arrangement is not apparent in the two-dimensional (2D) ϕ,ψ representation. In this article, we first represent the backbone of a protein as a pc-polyline that passes through the centers of its peptide planes. We then analyze the pc-polylines for six different sets of proteins with high quality crystal structures. The results show that SSE composition becomes recognizable in pc-polyline presentation and consequently the geometrical property of the pc-polyline of a protein could be used to assign its secondary structure. Furthermore, our analysis finds that for each of the six sets the total length of a pc-polyline increases linearly with the number of the peptide planes. Interestingly a comparison of the six regression lines shows that they have almost identical slopes but different intercepts. Most interestingly there exist decent linear correlations between the intercepts of the six lines and either the average helix contents or the average sheet contents and between the intercepts and the average backbone hydrogen bonding energetics. Finally, we discuss the implications of the identified correlations for structure classification and protein folding, and the potential applications of pc-polyline representation to structure prediction and protein design.  相似文献   

19.
The staphylococcal enterotoxin H (SEH; 217 aa, 25 kD) belongs to a family of superantigens that cause a massive immune response upon simultaneous binding to the T cell receptor (TCR) and the major histocompatibility complex class II. The SEH-TCR interaction is weak and amenable to studies using NMR methodology. Essentially, 2 mg of U{2H, 13C,15N}-labeled SEH was used for the complete sequential backbone assignment of SEH at 900 MHz. The protein secondary structure inferred from the chemical shift index (Cα and Cβ) is in very good agreement with the secondary structure elements of the X-ray structure.  相似文献   

20.
Self‐assembly of PAs composed of palmitic acid and several repeated heptad peptide sequences, C15H31CO‐(IEEYTKK)n‐NH2 (n = 1–4, represented by PA1–PA4), was investigated systematically. The secondary structures of the PAs were characterized by CD. PA3 and PA4 (n = 3 and 4, respectively) showed an α‐helical structure, whereas PA1 and PA2 (n = 1 and 2, respectively) did not display an α‐helical conformations under the tested conditions. The morphology of the self‐assembled peptides in aqueous medium was studied by transmission electron microscopy. As the number of heptad repeats in the PAs increased, the nanostructure of the self‐assembled peptides changed from nanofibers to nanovesicles. Changes of the secondary structures and the self‐assembly morphologies of PA3 and PA4 in aqueous medium with various cations were also studied. The critical micelle concentrations were determined using a pyrene fluorescence probe. In conclusion, this method may be used to design new peptide nanomaterials. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  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