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.
In proteins, the polypeptide chain forms a number of right-and left-handed helices and superhelices, right-and left-turned hairpins, and some other structures that are nonsuperimposable, although they are not mirror images of each other as the Lamino acids are not converted to the Damino acids. This property of protein structures will be referred to here as pseudo-chirality–or handedness. It has been shown that there are two kinds of handedness in proteins–helical handedness and handedness of arrangement. Some protein structures exhibit both the kinds of handedness. Handedness is observed at all levels of protein structural organization–from α-helices, β-strands, hairpins, βαβ-units up to complex structural motifs, superhelices, and supramolecular structures in fibrous and polymer proteins. There are several structures that have unique handedness in proteins, for example, α-helices, αα-corners, βαβ-units, abcd-units, and so on. This property of the polypeptide chain is of particular value in protein folding and protein modeling, because it drastically reduces the number of possible folds.  相似文献   

2.
The β-strand-α-helix-β-strand unit consists of two parallel, but not necessarily adjacent, β-strands which lie in a β-pleated sheet and are connected by one or more α-helices. This unit, which occurs in 17 functionally different globular proteins, may adopt a right- or a left-handed conformation. An analysis of the distribution shows that 57 out of the 58 units are right-handed. If the unit had no right-handed preference, the probability of observing such a distribution by chance is 10?16. This may be explained in terms of the twist of the β-sheet which is shown to favour a right-handed unit, as otherwise steric hindrance occurs in the loop regions. We show that the right-handed strand-helix-strand unit determines the sense of the super-secondary structure found in the dehydrogenases and of related folds found in other structures. The evolutionary relationships between proteins containing this unit are re-evaluated in terms of this preference. The high probability that the unit will fold with a right-handed conformation has implications for the prediction of tertiary structure.  相似文献   

3.
Infrared spectra of 2.5 mM solutions of β-lactoglobulin B were recorded as a function of pH (from pH 2 to pH 13) and as a function of temperature (from −100°C to +90°C). An analysis of the pH- and temperature-induced changes in the secondary structures was performed based on changes in the conformation-sensitive amide I bands of β-lactoglobulin. Whereas the total of β-structure remains constant (56–59%) between pH22 and pH 10, the proportions of the various β-components do change. In particular, the dimerization of the monomeric protein, induced by raising the pH from 20 to 3, leads to an increase in the intensity of the 1636 cm−1 band (associated with antiparallel β-sheet), at the expense of the 1626 cm−1 band (associated with exposed β-strands). Both the thermal and alkaline denaturation of β-lactoglobulin occur in two distinct stages. Although the spectra (i.e., the structures) after complete thermal or alkaline denaturation are clearly different, the spectrum of the protein after the first stage of thermal denaturation (at about 60°C) is the same as that after the first stage of alkaline denaturation (at pH 11), suggesting a common denaturation intermediate, which probably represents a crossover point in a complex potential hypersurface.  相似文献   

4.
Abstract

The crystal-state preferred conformations of two tripeptides, one tetrapeptide, and one pen- tapeptide, each containing a single residue of the chiral, Cα,α-disubstituted glycine Cα-methyl, Cα-benzylglycine [(αMe)Phe], have been determined by X-ray diffraction. The tripeptides are Z-L-(αMe)Phe-(Aib)2-OH dihydrate and Z-Aib-D-(αMe)Phe-Aib-OtBu, the tetrapeptide is Z-(Aib)2-D-(αMe)Phe-Aib-OtBu, and the pentapeptide is pBrBz-(Aib)2-DL-(αMe)Phe-(Aib)2-OtBu. While the two tripeptides are folded in a β-bend conformation, two such conformations are consecutively formed by the tetrapeptide. The pentapeptide adopts a regular 310-helix promoted by three consecutive β-bends. This study confirms the strong propensity of short peptides containing Cα-methylated α-aminoacids to fold into β-bends and 310-helical structures. Since Aib is achiral, the handedness of the observed bends and helices is dictated by the presence of the (αMe)Phe residue. In general, we have found that the relationship between (αMe)Phe chirality and helix handedness is opposite to that exhibited by protein aminoacids. A comparison with the preferred conformation of other extensively investigated Cα-methylated aminoacids is made.  相似文献   

5.
Cross-strand pair correlations are calculated for residue pairs in antiparallel β-sheet for two cases: pairs whose backbone atoms are hydrogen bonded together (H-bonded site) and pairs which are not (non-H-bonded site). The statistics show that this distinction is important. When glycine is located on the edge of a sheet, it shows a 3:1 preference for the H-bonded site. Thestrongest observed correlations are for pairs of disulfide-bonded cystines, many of which adopt a close-packed conformation with each cystine in a spiral conformation of opposite chirality to its partner. It is likely that these pairs are a signature for the family of small, cystine-rich proteins. Most other strong positive and negative correlations involve charged and polar residues. It appears that electrostatic compatibility is the strongest factor affecting pair correlation. Significant correlations are observed for β- and γ-branched residues inthe non-H-bonded site. An examination of the structures showsa directionality in side chain packing. There is a correlation between (1) the directionality in the packing interactions of non-H-bonded β- and γ-branched residue pairs, (2) the handedness of the observed enantiomers of chiral β-branched side chains, and (3) the handedness of the twist of β-sheet. These findings have implications for the formation of β-sheets during protein folding and the mechanism by which the sheet becomes twisted. © 1995 Wiley-Liss, Inc.  相似文献   

6.
The type and distribution of the β-helixes occurring in chloroform solutions of Boc-(L-Val-D-Val)6-OMe and Boc-(L-Val-D-Val)8-OMe have been studied by using 1H-nmr techniques. Right- and left-handed ↑↓β4.4-helices and left-handed β5.6-helices occur with the dodecapeptide. β4.4-Helices of opposite handedness occur also with the hexadecapeptide, but ↑↓β5.6-helices could not be detected with this oligomer. At equilibrium, at 25°C, the double helix of the dodecapeptide is only moderately populated. These results indicate that increasing the chain length has a destabilizing effect on the ↑↓β5.6-helices of D ,L -alternating oligovalines in chloroform solution.  相似文献   

7.
Leucine rich repeats (LRRs) are present in over 100,000 proteins from viruses to eukaryotes. The LRRs are 20–30 residues long and occur in tandem. LRRs form parallel stacks of short β-strands and then assume a super helical arrangement called a solenoid structure. Individual LRRs are separated into highly conserved segment (HCS) with the consensus of LxxLxLxxNxL and variable segment (VS). Eight classes have been recognized. Bacterial LRRs are short and characterized by two prolines in the VS; the consensus is xxLPxLPxx with Nine residues (N-subtype) and xxLPxxLPxx with Ten residues (T-subtype). Bacterial LRRs are contained in type III secretion system effectors such as YopM, IpaH3/9.8, SspH1/2, and SlrP from bacteria. Some LRRs in decorin, fribromodulin, TLR8/9, and FLRT2/3 from vertebrate also contain the motifs. In order to understand structural features of bacterial LRRs, we performed both secondary structures assignments using four programs—DSSP-PPII, PROSS, SEGNO, and XTLSSTR—and HELFIT analyses (calculating helix axis, pitch, radius, residues per turn, and handedness), based on the atomic coordinates of their crystal structures. The N-subtype VS adopts a left handed polyproline II helix (PPII) with four, five or six residues and a type I β-turn at the C-terminal side. Thus, the N-subtype is characterized by a super secondary structure consisting of a PPII and a β-turn. In contrast, the T-subtype VS prefers two separate PPIIs with two or three and two residues. The HELFIT analysis indicates that the type I β-turn is a right handed helix. The HELFIT analysis determines three unit vectors of the helix axes of PPII (P), β-turn (B), and LRR domain (A). Three structural parameters using these three helix axes are suggested to characterize the super secondary structure and the LRR domain.  相似文献   

8.
Gram-negative bacteria, such as Escherichia coli, frequently utilize tripartite efflux complexes belonging to the resistance-nodulation-division family to expel diverse toxic compounds from the cell. These systems contain a periplasmic membrane fusion protein (MFP) that is critical for substrate transport. We here present the x-ray structures of the CusB MFP from the copper/silver efflux system of E. coli. This is the first structure of any MFPs associated with heavy-metal efflux transporters. CusB bridges the inner-membrane efflux pump CusA and outer-membrane channel CusC to mediate resistance to Cu+ and Ag+ ions. Two distinct structures of the elongated molecules of CusB were found in the asymmetric unit of a single crystal, which suggests the flexible nature of this protein. Each protomer of CusB can be divided into four different domains, whereby the first three domains are mostly β-strands and the last domain adopts an entirely helical architecture. Unlike other known structures of MFPs, the α-helical domain of CusB is folded into a three-helix bundle. This three-helix bundle presumably interacts with the periplasmic domain of CusC. The N- and C-termini of CusB form the first β-strand domain, which is found to interact with the periplasmic domain of the CusA efflux pump. Atomic details of how this efflux protein binds Cu+ and Ag+ were revealed by the crystals of the CusB-Cu(I) and CusB-Ag(I) complexes. The structures indicate that CusB consists of multiple binding sites for these metal ions. These findings reveal novel structural features of an MFP in the resistance-nodulation-division efflux system and provide direct evidence that this protein specifically interacts with transported substrates.  相似文献   

9.
Type IV pili are bacterial extracellular filaments that can be retracted to create force and motility. Retraction is accomplished by the motor protein PilT. Crystal structures of Pseudomonas aeruginosa PilT with and without bound β,γ-methyleneadenosine-5′-triphosphate have been solved at 2.6 Å and 3.1 Å resolution, respectively, revealing an interlocking hexamer formed by the action of a crystallographic 2-fold symmetry operator on three subunits in the asymmetric unit and held together by extensive ionic interactions. The roles of two invariant carboxylates, Asp Box motif Glu163 and Walker B motif Glu204, have been assigned to Mg2+ binding and catalysis, respectively. The nucleotide ligands in each of the subunits in the asymmetric unit of the β,γ-methyleneadenosine-5′-triphosphate-bound PilT are not equally well ordered. Similarly, the three subunits in the asymmetric unit of both structures exhibit differing relative conformations of the two domains. The 12° and 20° domain rotations indicate motions that occur during the ATP-coupled mechanism of the disassembly of pili into membrane-localized pilin monomers. Integrating these observations, we propose a three-state “Ready, Active, Release” model for the action of PilT.  相似文献   

10.
Grid-free protein folding simulations based on sequence and secondary structure knowledge (using mostly experimentally determined secondary structure information but also analysing results from secondary structure predictions) were investigated using the genetic algorithm, a backbone representation, and standard dihedral angular conformations. Optimal structures are selected according to basic protein building principles. Having previously applied this approach to proteins with helical topology, we have now developed additional criteria and weights for β-strand- containing proteins, validated them on four small β-strand-rich proteins with different topologies, and tested the general performance of the method on many further examples from known protein structures with mixed secondary structural type and less than 100 amino acid residues.Topology predictions close to the observed experimental structures were obtained in four test cases together with fitness values that correlated with the similarity of the predicted topology to the observed structures. Root-mean-square deviation values of Cαatoms in the superposed predicted and observed structures, the latter of which had different topologies, were between 4.5 and 5.5 Å (2.9 to 5.1 Å without loops). Including 15 further protein examples with unique folds, root-mean-square deviation values ranged between 1.8 and 6.9 Å with loop regions and averaged 5.3 Å and 4.3 Å, including and excluding loop regions, respectively.  相似文献   

11.
Accumulating evidence suggests that soluble oligomers are more toxic than final fibrils of amyloid aggregations. Among the mixture of inter-converting intermediates with continuous distribution of sizes and secondary structures, oligomers in the β-barrel conformation – a common class of protein folds with a closed β-sheet – have been postulated as the toxic species with well-defined three-dimensional structures to perform pathological functions. A common mechanism for amyloid toxicity, therefore, implies that all amyloid peptides should be able to form β-barrel oligomers as the aggregation intermediates. Here, we applied all-atom discrete molecular dynamics (DMD) simulations to evaluate the formation of β-barrel oligomers and characterize their structures and dynamics in the aggregation of a seven-residue amyloid peptide, corresponding to the amyloid core of amyloid-β with a sequence of 16KLVFFAE22 (Aβ16-22). We carried out aggregation simulations with various numbers of peptides to study the size dependence of aggregation dynamics and assembly structures. Consistent with previous computational studies, we observed the formation of β-barrel oligomers in all-atom DMD simulations. Using a network-based approach to automatically identify β-barrel conformations, we systematically characterized β-barrels of various sizes. Our simulations revealed the conformational inter-conversion between β-barrels and double-layer β-sheets due to increased structural strains upon forming a closed β-barrel while maximizing backbone hydrogen bonds. The potential of mean force analysis further characterized the free energy barriers between these two states. The obtained structural and dynamic insights of β-barrel oligomers may help better understand the molecular mechanism of oligomer toxicities and design novel therapeutics targeting the toxic β-barrel oligomers. This article is part of a Special Issue entitled: Protein Aggregation and Misfolding at the Cell Membrane Interface edited by Ayyalusamy Ramamoorthy  相似文献   

12.
Zhang C  Kim SH 《Proteins》2000,40(3):409-419
The Greek key motifs are the topological signature of many beta-barrels and a majority of beta-sandwich structures. An updated survey of these structures integrates many early observations and newly emerging patterns and provides a better understanding of the unique role of Greek keys in protein structures. A stereotypical Greek key beta-barrel accommodates five or six strands and can have 12 possible topologies. All except one six-stranded topologies have been observed, and only one five-stranded topologies have been seen in actual structures. Of the representative beta-barrel structures analyzed here, half have left-handed Greek keys. This result challenges the empirical claim of the handedness regularity of Greek keys in beta-barrels. One of the five-stranded topologies that has not been observed in beta-barrels comprises two overlapping Greek keys. The two three-dimensional forms of this topology constitute a structural unit that is present in a vast majority of known beta-sandwich structures. Using this unit as the root, we have built a new taxonomy tree for the beta-sandwich folds and deduced a set of rules that appear to constrain how other beta-strands adjoin the unit to form a larger double-layered structure. These rules, though derived from a larger data set, are essentially the same as those drawn from earlier studies, suggesting that they may reflect the true topological constraints in the design of beta-sandwich structures. Finally, a novel variant of the Greek key motif (defined here as the twisted Greek key) has emerged which introduces loop crossings into the folded structures. Proteins 2000;40:409-419.  相似文献   

13.
The energies of two and three-chain antiparallel and parallel β-sheets have been minimized. The chains were considered to be equivalent. In each case, chains consisting of four and of eight l-alanine residues, respectively, with CH3CO- and -NHCH3 end groups were examined. Computations were carried out both for chains constrained to have a regular structure (i.e. the same φ and ψ dihedral angles for each residue) and for chains in which the regularity constraint was relaxed. All computed minimum-energy β-sheets were found to have a right-handed twist, as observed in proteins. As in the case of right-handed α-helices, it is the intrastrand non-bonded interaction energy that plays the key role in forcing β-sheets of l-amino acid residues to adopt a right-handed twist. The non-bonded energy contribution favoring the right-handed twist is the result of many small pairwise interatomic interactions involving the CβH3 groups. Polyglycine β-sheets, lacking the CβH3 side-chains, are not twisted. The twist of the poly-l-alanine sheet diminishes as the number of residues per chain increases, in agreement with observations. The twist of the four-residue chain increases somewhat (because of interstrand non-bonded interactions, also involving the CβH3 groups) in going from a single chain to a two-chain antiparallel structure, but then decreases slightly in going from a two-chain to a three-chain structure. β-Sheets in observed protein structures sometimes have a larger twist than those in the structures computed here. This may be due to irregularities in amino acid sequence and in hydrogenbonding patterns in the observed sheets, or to long-range interactions in proteins. The minimized energies of parallel β-sheets are considerably higher than those of the corresponding antiparallel β-sheets, indicating that parallel β-sheets are intrinsically less stable. This finding about the two kinds of β-sheets agrees with suggestions based on analyses of β-sheets observed in proteins. The energy difference between antiparallel and parallel β-sheets is due to closer packing of the chains and a more favorable alignment of the peptide dipoles in the antiparallel structures. The hydrogen-bond geometry in the computed antiparallel structures is very close to that proposed by Arnott et al. (1967) for the β-form of poly-l-alanine.  相似文献   

14.
Chiral nanostructures show macroscopic optical activity. Local optical activity and its handedness are not uniform in the nanostructure, and are spatially distributed depending on the shape of the nanostructure. In this study we fabricated curved chain nanostructures made of gold by connecting linearly two or more arc structures in a two‐dimensional plane. Spatial features of local optical activity in the chain structures were evaluated with near‐field circular dichroism (CD) imaging, and analyzed with the aid of classical electromagnetic simulation. The electromagnetic simulation predicted that local optical activity appears at inflection points where arc structures are connected. The handedness of the local optical activity was dependent on the handedness of the local chirality at the inflection point. Chiral chain structures have odd inflection points and the local optical activity distributed symmetrically with respect to structural centers. In contrast, achiral chain structures have even inflection points and showed antisymmetric distribution. In the near‐field CD images of fabricated chain nanostructures, the symmetric and antisymmetric distributions of local CD were observed for chiral and achiral chain structures, respectively, consistent with the simulated results. The handedness of the local optical activity was found to be determined by the handedness of the inflection point, for the fabricated chain structures having two or more inflection points. The local optical activity was thus governed primarily by the local chirality of the inflection points for the gold chain structures. The total effect of all the inflection points in the chain structure is considered to be a predominant factor that determines the macroscopic optical activity. Chirality 28:540–544, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

15.
The behavior of 1H-nmr spin–lattice relaxation times as a function of frequency is examined for powders of α- and β-cyclodextrins, dextran B512F, and their “deuterated” analogues, where deuterium replaces exchangeable protons. A crossover from two- to one-component relaxation decays is observed with increasing frequency (ν). By considering the effect of spin diffusion in spin–lattice relaxation measurements as a function of the frequency, estimates of domain sizes (~ 7–10 Å) consistent with the lower limit of spatial resolution of the Goldman–Shen experiments are obtained.  相似文献   

16.
A computer-assisted procedure, based upon a branch of mathematics known as graph theory, has been developed to recognize secondary structure elements in proteins from their corresponding nuclear Overhauser effect spectroscopy (NOESY)-type spectra and to carry out their sequential assignment. In the method, NOE connectivity templates characteristic of regular secondary structures are identified in the spectra. Resonance assignment is then achieved by connecting these NOE patterns of secondary structure together, and thereby matching connected spin systems to specific parts of the primary sequence. The range of NOE-graph templates of secondary structure motifs, incorporating α-helices and β-strand motifs, has been examined for reliability and extent of secondary structure identification in a data base composed of the high resolution structures of 20 proteins. The analysis identified several robust NOE-graph templates and supports the implementation of an ordered search strategy. The method, known as SERENDIPITY, has been applied to the analysis of nuclear Overhauser effect data from a three-dimensional time-shared nuclear Overhauser effect spectroscopy (13C, 15N) heteronuclear single quantum correlation spectrum of the (α + β) type protein HU from Bacillus stearothermophilus. The arrangement of the elucidated elements of secondary structure is very similar to that of the x-ray and nmr structures of HU. In addition, our analysis revealed a pattern of interstrand nuclear Overhauser effect in the β-arm region (residues 53–76) of HU, which suggest irregularities, not reported in the x-ray structure, in both strands of the β-arm at Ala57 and Pro72, respectively. At these residues, both strands of the β-arm appear to flip inside out before continuing as a regular antiparallel β-sheet. © 1996 John Wiley & Sons, Inc.  相似文献   

17.
On the origin of biological chirality via natural beta-decay   总被引:1,自引:0,他引:1  
Experimental evidence that longitudinally polarized electrons having the handedness characteristic of terrestrial beta-decay electrons preferentially remove D-leucine from a racemic mixture, coupled with the probable presence of14C in pre-biotic molecules, offers a plausible hypothesis for the origin of biomolecular handedness.  相似文献   

18.
In fully developed androgen-induced hypertrophy of female mouse kidney, β-adrenergic receptors per unit membrane protein were increased approx. 2.5-fold, as measured by the binding of [125I]iodocyanopindolol, with no change in apparent dissociation constants (Kd range 20–25 pM). Membrane protein relative to total kidney protein, Na+/K+-dependent ATPase (EC 3.6.1.3) and 5′-nucleotidase (EC 3.1.3.5) activities and cholesterol content per unit membrane protein did not differ significantly in preparations from control and treated animals. The binding of iodocyanopindolol to kidney membranes was characterized with respect to association and dissociation kinetics, and also in regard to the less-specific contributions of other major catecholamine or indolamine receptors, using mixtures of the corresponding specific competitors. β1-selective drugs, practolol and metoprolol, and β2-selective agents, IPS-339 and zinterol, were competed with iodocyanopindolol to assess the receptor type specificity, and the ensuing binding profiles were dissected by a nonlinear regression analysis as described by Munson, P.J. and Rodbard, D. (Anal. Biochem. (1982) 107, 220–239). Most of the androgen-induced β-adrenergic receptors had the binding properties corresponding to β2-subtype. No consistent increase in the density of β1-adrenergic recepotors could be shown.  相似文献   

19.
G‐rich sequences can adopt four‐stranded helical structures, called G‐quadruplexes, that self‐assemble around monovalent cations like sodium (Na+) and potassium (K+). Whether similar structures can be formed from xeno‐nucleic acid (XNA) polymers with a shorter backbone repeat unit is an unanswered question with significant implications on the fold space of functional XNA polymers. Here, we examine the potential for TNA (α‐l ‐threofuranosyl nucleic acid) to adopt a four‐stranded helical structure based on a planar G‐quartet motif. Using native polyacrylamide gel electrophoresis (PAGE), circular dichroism (CD) and solution‐state nuclear magnetic resonance (NMR) spectroscopy, we show that despite a backbone repeat unit that is one atom shorter than the backbone repeat unit found in DNA and RNA, TNA can self‐assemble into stable G‐quadruplex structures that are similar in thermal stability to equivalent DNA structures. However, unlike DNA, TNA does not appear to discriminate between Na+ and K+ ions, as G‐quadruplex structures form equally well in the presence of either ion. Together, these findings demonstrate that despite a shorter backbone repeat unit, TNA is capable of self‐assembling into stable G‐quadruplex structures.  相似文献   

20.
A unified coarse-grained model of three major classes of biological molecules—proteins, nucleic acids, and polysaccharides—has been developed. It is based on the observations that the repeated units of biopolymers (peptide groups, nucleic acid bases, sugar rings) are highly polar and their charge distributions can be represented crudely as point multipoles. The model is an extension of the united residue (UNRES) coarse-grained model of proteins developed previously in our laboratory. The respective force fields are defined as the potentials of mean force of biomacromolecules immersed in water, where all degrees of freedom not considered in the model have been averaged out. Reducing the representation to one center per polar interaction site leads to the representation of average site–site interactions as mean-field dipole–dipole interactions. Further expansion of the potentials of mean force of biopolymer chains into Kubo’s cluster-cumulant series leads to the appearance of mean-field dipole–dipole interactions, averaged in the context of local interactions within a biopolymer unit. These mean-field interactions account for the formation of regular structures encountered in biomacromolecules, e.g., α-helices and β-sheets in proteins, double helices in nucleic acids, and helicoidally packed structures in polysaccharides, which enables us to use a greatly reduced number of interacting sites without sacrificing the ability to reproduce the correct architecture. This reduction results in an extension of the simulation timescale by more than four orders of magnitude compared to the all-atom representation. Examples of the performance of the model are presented.
Figure
Components of the Unified Coarse Grained Model (UCGM) of biological macromolecules  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