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.
The nest is a protein motif of three consecutive amino acid residues with dihedral angles 1,2‐αRαL (RL nests) or 1,2‐αLαR (LR nests). Many nests form a depression in which an anion or δ‐negative acceptor atom is bound by hydrogen bonds from the main chain NH groups. We have determined the extent and nature of this bridging in a database of protein structures using a computer program written for the purpose. Acceptor anions are bound by a pair of bridging hydrogen bonds in 40% of RL nests and 20% of LR nests. Two thirds of the bridges are between the NH groups at Positions 1 and 3 of the motif (N1N3‐bridging)—which confers a concavity to the nest; one third are of the N2N3 type—which does not. In bridged LR nests N2N3‐bridging predominates (14% N1N3: 75% N2N3), whereas in bridged RL nests the reverse is true (69% N1N3: 25% N2N3). Most bridged nests occur within larger motifs: 45% in (hexapeptide) Schellman loops with an additional 4 → 0 hydrogen bond (N1N3), 11% in Schellman loops with an additional 5 → 1 hydrogen bond (N2N3), 12% in a composite structure including a type 1β‐bulge loop and an asx‐ or ST‐ motif (N1N3)—remarkably homologous to the N1N3‐bridged Schellman loop—and 3% in a composite structure including a type 2β‐bulge loop and an asx‐motif (N2N3). A third hydrogen bond is a previously unrecognized feature of Schellman loops as those lacking bridged nests have an additional 4 → 0 hydrogen bond. Proteins 2014; 82:3023–3031. © 2014 Wiley Periodicals, Inc.  相似文献   

2.
The concept of bends or chain reversals [nonhelical dipeptide sequences in which the distance R3 (i,i+3) between the Cα atoms of residues i and i+3 is ≦ 7.0 Å] has been extended to define double bends as tripeptide sequences, not in an α-helix, in which two successive distances R3(i,i+3) and R3 (i+1, i+4) are both ≦7.0 Å, with analogous definitions for higher-order multiple bends. A sample of 23 proteins, consisting of 4050 residues, contains 235 single, 58 double, and 11 higher-order multiple bends. Multiple bends may occur as combinations of the “standard” type I, II, and III chain reversals (as well as their mirror images), but usually they require distortions from these well-defined conformations. The frequency of occurrence of amino acids often differs significantly between single and multiple bends. The probability distribution of R3 distances does not differ in single and multiple bends. However, R4 (the distance between the Cα atoms of residues i and i+4) in multiple bends is generally shorter than in tripeptide sequences containing single bends. The value of R4 in many multiple bends is near those for α-helices. In some other multiple bends, R4 is even shorter, indicating that these structures are very compact. The signs of the dihedral angles about the virtual bonds connecting Cα atoms and the values of curvature and torsion, as defined by means of differential geometry, indicate that there is a preference for single and multiple bends to be right-handed (like an α-helical sequence, for example) and that there is a strong tendency to conserve the handedness in both single-bend components of many multiple bends. These often have a strong resemblance to distorted single turns of an α-helix and do not constitute chain reversals. Double bends, in which the signs of two successive virtual-bond dihedral angles differ, have conformations that are very different from an α-helix. They act as chain reversals occuring over three residues. These chain reversals have not been described previously. Multiple bends may play an important role in protein folding because they occur fairly frequently in proteins and cause major changes in the direction of the polypeptide chain.  相似文献   

3.
Despite recent improvements in computational methods for protein design, we still lack a quantitative, predictive understanding of the intrinsic probabilities for amino acids to adopt particular side‐chain conformations. Surprisingly, this question has remained unsettled for many years, in part because of inconsistent results from different experimental approaches. To explicitly determine the relative populations of different side‐chain dihedral angles, we performed all‐atom hard‐sphere Langevin Dynamics simulations of leucine (Leu) and isoleucine (Ile) dipeptide mimetics with stereo‐chemical constraints and repulsive‐only steric interactions between non‐bonded atoms. We determine the relative populations of the different χ1 and χ2 dihedral angle combinations as a function of the backbone dihedral angles ? and ψ. We also propose, and test, a mechanism for inter‐conversion between the different side‐chain conformations. Specifically, we discover that some of the transitions between side‐chain dihedral angle combinations are very frequent, whereas others are orders of magnitude less frequent, because they require rare coordinated motions to avoid steric clashes. For example, to transition between different values of χ2, the Leu side‐chain bond angles κ1 and κ2 must increase, whereas to transition in χ1, the Ile bond angles λ1 and λ2 must increase. These results emphasize the importance of computational approaches in stimulating further experimental studies of the conformations of side‐chains in proteins. Moreover, our studies emphasize the power of simple steric models to inform our understanding of protein structure, dynamics, and design. Proteins 2015; 83:1488–1499. © 2015 Wiley Periodicals, Inc.  相似文献   

4.
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   

5.
6.
The recently determined C. elegans P‐glycoprotein (Pgp) structure revealed significant deviations compared to the original mouse Pgp structure, which suggested possible misinterpretations in the latter model. To address this concern, we generated an experimental electron density map from single‐wavelength anomalous dispersion phasing of an original mouse Pgp dataset to 3.8 Å resolution. The map exhibited significantly more detail compared to the original MAD map and revealed several regions of the structure that required de novo model building. The improved drug‐free structure was refined to 3.8 Å resolution with a 9.4 and 8.1% decrease in Rwork and Rfree, respectively, (Rwork = 21.2%, Rfree = 26.6%) and a significant improvement in protein geometry. The improved mouse Pgp model contains ~95% of residues in the favorable Ramachandran region compared to only 57% for the original model. The registry of six transmembrane helices was corrected, revealing amino acid residues involved in drug binding that were previously unrecognized. Registry shifts (rotations and translations) for three transmembrane (TM)4 and TM5 and the addition of three N‐terminal residues were necessary, and were validated with new mercury labeling and anomalous Fourier density. The corrected position of TM4, which forms the frame of a portal for drug entry, had backbone atoms shifted >6 Å from their original positions. The drug translocation pathway of mouse Pgp is 96% identical to human Pgp and is enriched in aromatic residues that likely play a collective role in allowing a high degree of polyspecific substrate recognition.  相似文献   

7.
The side‐chain dihedral angle distributions of all amino acids have been measured from myriad high‐resolution protein crystal structures. However, we do not yet know the dominant interactions that determine these distributions. Here, we explore to what extent the defining features of the side‐chain dihedral angle distributions of different amino acids can be captured by a simple physical model. We find that a hard‐sphere model for a dipeptide mimetic that includes only steric interactions plus stereochemical constraints is able to recapitulate the key features of the back‐bone dependent observed amino acid side‐chain dihedral angle distributions of Ser, Cys, Thr, Val, Ile, Leu, Phe, Tyr, and Trp. We find that for certain amino acids, performing the calculations with the amino acid of interest in the central position of a short α‐helical segment improves the match between the predicted and observed distributions. We also identify the atomic interactions that give rise to the differences between the predicted distributions for the hard‐sphere model of the dipeptide and that of the α‐helical segment. Finally, we point out a case where the hard‐sphere plus stereochemical constraint model is insufficient to recapitulate the observed side‐chain dihedral angle distribution, namely the distribution P3) for Met. Proteins 2014; 82:2574–2584. © 2014 Wiley Periodicals, Inc.  相似文献   

8.
Molecular transduction of biological signals is understood primarily in terms of the cooperative structural transitions of protein macromolecules, providing a mechanism through which discrete local structure perturbations affect global macromolecular properties. The recognition that proteins lacking tertiary stability, commonly referred to as intrinsically disordered proteins (IDPs), mediate key signaling pathways suggests that protein structures without cooperative intramolecular interactions may also have the ability to couple local and global structure changes. Presented here are results from experiments that measured and tested the ability of disordered proteins to couple local changes in structure to global changes in structure. Using the intrinsically disordered N‐terminal region of the p53 protein as an experimental model, a set of proline (PRO) and alanine (ALA) to glycine (GLY) substitution variants were designed to modulate backbone conformational propensities without introducing non‐native intramolecular interactions. The hydrodynamic radius (Rh) was used to monitor changes in global structure. Circular dichroism spectroscopy showed that the GLY substitutions decreased polyproline II (PPII) propensities relative to the wild type, as expected, and fluorescence methods indicated that substitution‐induced changes in Rh were not associated with folding. The experiments showed that changes in local PPII structure cause changes in Rh that are variable and that depend on the intrinsic chain propensities of PRO and ALA residues, demonstrating a mechanism for coupling local and global structure changes. Molecular simulations that model our results were used to extend the analysis to other proteins and illustrate the generality of the observed PRO and alanine effects on the structures of IDPs. Proteins 2014; 82:3373–3384. © 2014 Wiley Periodicals, Inc.  相似文献   

9.
Sumana Giddu  Fei Xu  Vikas Nanda 《Proteins》2013,81(3):386-393
Stability of the collagen triple helix is largely governed by its imino acid content, namely the occurrence of proline and 4R‐hydroxyproline at the X and Y positions, respectively, of the periodic (Gly‐X‐Y)n sequence. Although other amino acids at these positions reduce stability of the triple helix, this can be partially compensated by introducing intermolecular side‐chain salt bridges. This approach was previously used to design an abc‐type heterotrimer composed of one basic, one acidic, and one neutral imino acid rich chain (Gauba and Hartgerink, J Am Chem Soc 2007;129:15034–15041). In this study, an abc‐type heterotrimer was designed to be the most stable species using a sequence recombination strategy that preserved both the amino acid composition and the network of interchain salt bridges of the original design. The target heterotrimer had the highest Tm of 50°C, 7°C greater than the next most stable species. Stability of the heterotrimer decreased with increasing ionic strength, consistent with the role of intermolecular salt bridges in promoting stability. Quantitative meta‐analysis of these results and published stability measurements on closely related peptides was used to discriminate the contributions of backbone propensity and side‐chain electrostatics to collagen stability. Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

10.
Human factor XI (hFXI) is a 160‐kDa disulphide‐linked homodimer zymogen involved in the coagulation cascade. Its deficiency results in bleeding diathesis referred to as hemophilia C. hFXI bears five N‐glycosylation consensus sites per monomer, N72, N108, N335 on the heavy chain and N432, N473 on the light chain. This study reports the first in‐depth glycosylation analysis of hFXI based on advanced MS approaches. Hydrophilic interaction LC and MS characterization and quantification of the N‐glycans showed that the two major forms are complex biantennary mono‐α2,6‐sialylated (A2S1, 20%) and bis‐α2,6‐sialylated structures (A2S2, 66%). Minor triantennary structures (A3S3F, ~1.5%; A3S3, ~2%) were also identified. MS analyses of intact hFXI revealed full occupation of two of the three heavy‐chain glycosites and almost full‐site occupancy of the light chain. Analysis of hFXI glycopeptides by LC‐MS/MS enabled site‐specific glycan profiling and occupancy. It was evidenced that N335 was not glycosylated and that N72 and N108 were fully occupied, whereas N432 and N473 were occupied at about 92 and 95%, respectively. We also identified a new glycosite of the noncanonical format NXC at N145, occupied at around 5%. These data provide valuable structural information useful to understand the potential roles of N‐glycosylation on hFXI function and could serve as a structural reference.  相似文献   

11.
12.
The crystal structure of a tripeptide Boc‐Leu‐Val‐Ac12c‐OMe ( 1 ) is determined, which incorporates a bulky 1‐aminocyclododecane‐1‐carboxylic acid (Ac12c) side chain. The peptide adopts a semi‐extended backbone conformation for Leu and Val residues, while the backbone torsion angles of the Cα,α‐dialkylated residue Ac12c are in the helical region of the Ramachandran map. The molecular packing of 1 revealed a unique supramolecular twisted parallel β‐sheet coiling into a helical architecture in crystals, with the bulky hydrophobic Ac12c side chains projecting outward the helical column. This arrangement resembles the packing of peptide helices in crystal structures. Although short oligopeptides often assemble as parallel or anti‐parallel β‐sheet in crystals, twisted or helical β‐sheet formation has been observed in a few examples of dipeptide crystal structures. Peptide 1 presents the first example of a tripeptide showing twisted β‐sheet assembly in crystals. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
In this work, a new asymmetrical backbone thienobenzodithiophene (TBD) containing four aromatic rings is designed, and then four polymers PTBD‐BZ, PTBD‐BDD, PTBD‐FBT, and PTBD‐Tz are synthesized. The planar and high degree of π‐conjugation configuration can guarantee effective charge carrier transport and the distinct flanked dihedral angles between the TBD core and conjugated side chain can subtly regulate the molecular aggregation and crystallinity. The four polymer/3,9‐bis(2‐methylene‐(3‐(1,1‐dicyanomethylene)‐indanone)‐5,5,11,11‐tetrakis(4‐hexylphenyl)‐dithieno[2,3‐d:2′,3′‐d′]‐s‐indaceno[1,2‐b:5,6‐b′]‐dithiophene (ITIC) blending films exhibit predominantly face‐on orientation. The photovoltaic devices based on wide bandgap polymers PTBD‐BZ and PTBD‐BDD achieve power conversion efficiencies (PCEs) as high as 12.02% and 11.39% without any post‐treatment. For the medium bandgap polymers PTBD‐FBT and PTBD‐Tz, the devices also show good PCEs of 10.18% and 11.02% with high VOC of 0.94 and 1.02 V, respectively, which indicates simultaneously achieving a VOC > 1 V and a high JSC is feasible to further improve the PSCs' performance by modifying this new backbone. This work reveals that the versatile asymmetric backbone is an excellent moiety to construct light‐harvesting copolymers and to modulate the microstructure for highly efficient PSCs.  相似文献   

14.
Dihedral conformational transitions are analyzed systematically in a model globular protein, cytochrome P450cam, to examine their structural and chemical dependences through combined conventional molecular dynamics (cMD), accelerated molecular dynamics (aMD) and adaptive biasing force (ABF) simulations. The aMD simulations are performed at two acceleration levels, using dihedral and dual boost, respectively. In comparison with cMD, aMD samples protein dihedral transitions approximately two times faster on average using dihedral boost, and ~3.5 times faster using dual boost. In the protein backbone, significantly higher dihedral transition rates are observed in the bend, coil, and turn flexible regions, followed by the β bridge and β sheet, and then the helices. Moreover, protein side chains of greater length exhibit higher transition rates on average in the aMD‐enhanced sampling. Side chains of the same length (particularly Nχ = 2) exhibit decreasing transition rates with residues when going from hydrophobic to polar, then charged and aromatic chemical types. The reduction of dihedral transition rates is found to be correlated with increasing energy barriers as identified through ABF free energy calculations. These general trends of dihedral conformational transitions provide important insights into the hierarchical dynamics and complex free energy landscapes of functional proteins. Proteins 2016; 84:501–514. © 2016 Wiley Periodicals, Inc.  相似文献   

15.
Maurotoxin (MTX) is a 34‐residue toxin that was isolated initially from the venom of the scorpion Scorpio maurus palmatus. Unlike the other toxins of the α‐KTx6 family (Pi1, Pi4, Pi7, and HsTx1), MTX exhibits a unique disulfide bridge organization of the type C1? C5, C2? C6, C3? C4, and C7? C8 (instead of the conventional C1? C5, C2? C6, C3? C7, and C4? C8, herein referred to as Pi1‐like) that does not prevent its folding along the classic α/β scaffold of scorpion toxins. MTXPi1 is an MTX variant with a conventional pattern of disulfide bridging without any primary structure alteration of the toxin. Here, using MTX and/or MTXPi1 as models, we investigated how the type of folding influences toxin recognition of the Shaker B potassium channel. Amino acid residues of MTX that were studied for Shaker B recognition were selected on the basis of their homologous position in charybdotoxin, a three disulfide‐bridged scorpion toxin also active on this channel type. These residues favored either an MTX‐ or MTXPi1‐like folding. Our data indicate clearly that Lys23 and Tyr32 (two out of ten amino acid residues studied) are the most important residues for Shaker B channel blockage by MTX. For activity on SKCa channels, the same amino acid residues also affect, directly or indirectly, the recognition of SK channels. The molecular modeling technique and computed docking indicate the existence of a correlation between the half cystine pairings of the mutated analogs and their activity on the Shaker B K+ channel. Overall, mutations in MTX could, or could not, change the reorganization of disulfide bridges of this molecule without affecting its α/β scaffold. However, changing of the peptide backbone (cross linking disulfide bridges from MTX‐like type vs MTXPi1‐like type) appears to have less impact on the molecule activity than mutation of certain key amino acids such as Lys23 and Tyr32 in this toxin. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
A 1:2 complex of hexamethylenetetramine with mercury(II) thiocyanate, (CH2)6N4·2Hg(SCN)2, was prepared and shown by X-ray crystallography to be polymeric. The mixed-ligand complex crystallizes in the space group P21/m, with a = 6.059(2), b = 19.710(5), c = 7.895(2) Å, β = 105.63(2)°, and Z = 2. The structure was refined to RF = 0.060 for 1634 observed MoKα diffractometer data. Mercury(II) atoms in a row are linked pairwise by two thiocyanato groups in an end-to-end bridging mode, to give an infinite chain running in the a direction. Two neighboring chains are further laterally connected, successively by bidentate organic ligands which lie on a crystallographic mirror plane. The coordination geometry about Hg(II) is distorted tetragonal pyramidal, the metal atom binding strongly to two S atoms and a tertiary amino N atom (apex), and weakly to two thiocyanato N atoms.  相似文献   

17.
A new graph–theoretical approach called thermodynamic sampling of amino acid residues (TSAR) has been elaborated to explicitly account for the protein side chain flexibility in modeling conformation‐dependent protein properties. In TSAR, a protein is viewed as a graph whose nodes correspond to structurally independent groups and whose edges connect the interacting groups. Each node has its set of states describing conformation and ionization of the group, and each edge is assigned an array of pairwise interaction potentials between the adjacent groups. By treating the obtained graph as a belief‐network—a well‐established mathematical abstraction—the partition function of each node is found. In the current work we used TSAR to calculate partition functions of the ionized forms of protein residues. A simplified version of a semi‐empirical molecular mechanical scoring function, borrowed from our Lead Finder docking software, was used for energy calculations. The accuracy of the resulting model was validated on a set of 486 experimentally determined pKa values of protein residues. The average correlation coefficient (R) between calculated and experimental pKa values was 0.80, ranging from 0.95 (for Tyr) to 0.61 (for Lys). It appeared that the hydrogen bond interactions and the exhaustiveness of side chain sampling made the most significant contribution to the accuracy of pKa calculations. Proteins 2011; © 2011 Wiley‐Liss, Inc.  相似文献   

18.
The protein tyrosine phosphatase SHP‐1 plays an important role in many physiological and pathophysiological processes. This phosphatase is activated through binding of ligands to its SH2‐domains, mainly to the N‐terminal one. Based on a theoretical docking model, backbone‐to‐side chain cyclized octapeptides were designed as ligands. Assembly of such modelled structures required the synthesis of N‐functionalized tyrosine derivatives and their incorporation into the sequence. Because of difficulties encountered in the condensation of N‐protected amino acids to the N‐alkylated tyrosine‐peptide we synthesized and used preformed dipeptide building units. As all attempts to obtain phosphorylated dipeptide units failed, the syntheses had to be performed with a free phenolic function. Use of different N‐alkyl or cycloalkyl residues in the N‐functionalized side chains allowed to investigate the effect of ring size, flexibility and hydrophobicity of formed lactam bridges on stimulatory activity. All tested linear and cyclic octapeptides stimulate the phosphatase activity of SHP‐1. Stimulatory activities of cyclic ligands increase with the chain length of the lactam bridges resulting in increased flexibility and better entropic preformation of the binding conformation. The strong activity of some cyclic octapeptides supports the modelled structure. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
20.
The three-dimensional structure of native SHL-I, a lectin from the venom of the Chinese bird spider Selenocosmia huwena, has been determined from two-dimensional 1H NMR spectroscopy recorded at 500 and 600 MHz. The best 10 structures have NOE violation <0.3 Å, dihedral violation <2 deg, and average root-mean-square differences of 0.85 + 0.06 Å over backbone atoms. The structure consists of a three-stranded antiparallel -sheet and three turns. The three disulfide bridges and three-stranded antiparallel -sheet form a inhibitor cystine knot motif which is adopted by several other small proteins, such as huwentoxin-I, -conotoxin, and gurmarin. The C-terminal fragment from Leu28 to Trp32 adopts two sets of conformations corresponding to the cis and trans conformations of Pro31. The structure of SHL-I also has high similarity with that of the N-terminus of hevein, a lectin from rubber-tree latex.  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