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 15 毫秒
1.
The principal elements of the secondary, tertiary and quaternary structure of the tetrameric human plasma prealbumin molecule have been determined by Fourier refinement of X-ray diffraction data at 1.8 Å resolution. The subunit has an extensive β-structure composed of eight strands organised into two four-stranded sheets. There is also one short α-helix. The tertiary structure is largely determined by the association of the two β-sheets. Important contributions to the tertiary structure are made by three tyrosines and one aspartic acid involved in side-chain-main-chain interactions; a buried histidine associated with a group of internal water molecules; and a compact cluster of seven aromatic residues. Quaternary interactions occur at two sets of interfaces closely organised around two of the three molecular 2-fold axes. The exclusive monomer-monomer interface is chiefly concerned with antiparallel hydrogen bond interactions which extend the two four-stranded sheets in the monomers to eight-stranded sheets in the dimer. One of the sheet interactions includes water molecules and tyrosine hydroxyls in the hydrogen bond pattern. The dimers associate through both hydrophilic and hydrophobic interactions at interfaces that involve all four monomers.  相似文献   

2.
3.
The conformation of the globular dimer (G2), the tailed asymmetric dodecamer (A12, also containing some tailed octamer A8) and the globular tetramer (G4, prepared by removing the collagen-like tail from A12) of acetylcholinesterase (acetylcholine acetylhydrolase, EC 3.1.1.7) was studied by circular dichroism (CD) in the ultraviolet region. The G2 and G4 forms had similar conformation with about 40% α-helix, 35% β-sheets and 4% β-turns; the tailed form had a lower helicity (about 34%) and β-form (about 25%) content probably because of the presence of the tail whose CD spectrum resembles that of an unordered form, but it had about the same amount of β-turns as the other two forms. All three forms also had similar CD spectra in the near-ultraviolet region due to their non-peptide chromophores. The pH, thermal and urea denaturation of the three acetylcholinesterase forms was also similar to each other. The pH-dependency of both the enzymatic activity and CD intensity of the three forms showed bell-shaped curves with a plateau at pH 7–8. The activity was completely lost at pH below 5 or above 10, but the corresponding CD spectra retained 70–80% of the original magnitudes. Thermal denaturation of the three forms at pH 7.5 showed a conformational transition and loss of activity between 30 and 40°C, but the CD intensity of the helical band at 222 nm was reduced by only 20–30%. Urea denaturation of the three form began at 1 M urea; it was protein concentration- and time-dependent. Again, the activity disappeared faster than the decreasing CD intensity. Thus, the overall conformation of the three acetylcholinesterase forms appears to be relatively stable, but their active site is easily perturbed by changing the environment. The loss of activity correlated well with the disapperance of the CD band of tryptophan(s) in the near-ultraviolet region, suggesting that the Trp residue(s) might be at or near the active center of the enzyme.  相似文献   

4.
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   

5.
The pectate lyases, PelC and PelE, have an unusual folding motif, known as a parallel β-helix, in which the polypeptide chain is coiled into a larger helix composed of three parallel β-sheets connected by loops having variable lengths and conformations. Since the regular secondary structure consists almost entirely of parallel β-sheets these proteins provide a unique opportunity to study the effect of parallel β-helical structure on circular dichroism (CD). We report here the CD spectra of PelC and PelE in the presence and absence of Ca2+, derive the parallel β-helical components of the spectra, and compare these results with previous CD studies of parallel β-sheet structure. The shape and intensity of the parallel β-sheet spectrum is distinctive and may be useful in identifying other proteins that contain the parallel β-helical folding motif. © 1995 Wiley-Liss, Inc.  相似文献   

6.
Beta-turns in proteins   总被引:40,自引:0,他引:40  
The X-ray atomic co-ordinates from 29 proteins of known sequence and structure were utilized to elucidate 459 β-turns in regions of chain reversals. Tetrapeptides whose αCiαC(i + 3) distances were below 7 Å and not in a helical region were characterized as β-turns. In addition, β-turns were considered to have hydrogen bonding if their computed O(i)N(i + 3) distances were ≤3.5 Å. The torsion angles of 26 proteins containing 421 β-turns were examined and classified into 11 bend types based on the (φ, ψ) dihedral angles of the i + 1 and i + 2 bend residues. The average frequency of β-turns is 32% as compared to the 38% helices and 20% β-sheets in the 29 proteins. The most frequently occurring bend residues are Asn, Cys, Asp in the first position, Pro, Ser, Lys in the second position, Asn, Asp, Gly in the third position, and Trp, Gly, Tyr in the fourth position. Residues with the highest β-turn potential in all four positions are Pro, Gly, Asn, Asp, and Ser with the most hydrophobic residues (i.e. Val, IIe, and Leu) showing the lowest bend potential. However, in the region just beyond the β-turns, hydrophobic residues occur with greater frequency than do hydrophilic residues. An environmental analysis of β-turn neighboring residues shows that reverse chain folding is stabilized by anti-parallel β-sheets as well as helix-helix and α-β interactions. The β-turn potential at the 12 positions adjacent to and including the bend were plotted for the 20 amino acids and showed dramatic positional preferences, which may be classified according to the nature of the side-chains. An examination of the 27 β-turns in elastase showed that 21 were found in identical positions as those in α-chymotrypsin. However, only 37 of the 84 bend residues were conserved, indicating that structural similarity may persist despite differences in sequence homology. A survey of residues occupying bend types I′, II′ and III′ showed that Gly appeared most frequently in the third position in bend types I′ and III′ as well as in the second position in bend types II′ and III′. Fourteen hydrogenbonded type II bends were found without a Gly at the third position, contrary to the energy calculations. Eight type VI bends with a cis Pro at the third position were also elucidated.  相似文献   

7.
A method to identify β-sheets in globular proteins from extended strands, using only α-carbon positions, has been developed. The strands that form β-sheets are picked up by means of simple distance criteria. The method has been tested by applying it to three proteins with accurately known secondary structures. It has also been applied to ten other proteins wherein only α-carbon coordinates are available, and the list of β-sheets obtained. The following points are worth noting: (i) The sheets identified by the algorithm are found to agree satisfactorily with the reported ones based on backbone hydrogen bonding, wherever this information is available. (ii) β-Strands that do not form parts of any sheet are a common feature of protein structures. (iii) Such isolated β-strands tend to be short. (iv) The conformation corresponding to the preferred right-handed twist of the sheet is overwhelmingly observed in both the sheet-forming and isolated β-strands.  相似文献   

8.
The present work treats the conformational and geometrical properties of antiparallel and mixed β-sheets, following the approach described in the first paper of this series (Salemme & Weatherford, 1981). As shown, antiparallel structures possess conformational degrees of freedom that allow them to assume a greater diversity of spatial configurations than occur in parallel sheets. This configurational flexibility principally owes its origin to the potential variability of the interchain hydrogen bond geometry in antiparallel structures. As a consequence of this inherent elasticity of antiparallel β-sheets, their extended spatial configurations strongly reflect effects arising from local or extended side-chain packing interactions. Although the continuously deformable characteristics of antiparallel sheets frequently result in the attainment of spatial configurations that cannot be quantitatively modeled as hydrogen-bonded arrays of conformationally identical polypeptide chains, several observed sheets are nevertheless shown to be accurately approximated as conformationally regular arrays that typically yield model versus observed fits of less than 1 Å for corresponding α-carbon atoms.  相似文献   

9.
Several globular proteins have values of the Scheraga–Mandelkern β parameter significantly below the theoretical minimum value, β0 = 2.112 × 106, for an impermeable sphere. Using the Felderhof–Deutch generalization of the Debye–Bueche–Brinkman theory of hydrodynamics of porous spheres, we have shown that values of β slightly below this supposed minimum are theoretically expected. A porous sphere of uniform density has a minimum β of 2.084 × 106 at a Debye shielding ratio of 6.5, corresponding, for example, to a sphere radius of 11 Å and an inverse hydrodynamic shielding length of 0.6 Å?1, values not far from those of small proteins. A two-layer porous sphere model gives similar results. Although this is the first theoretical explanation of values of β below β0, the theory is incomplete since β values as low as 2.03 × 106 are observed.  相似文献   

10.
Plastocyanin and azurin form a family of small copper-containing proteins, active in the electron transport systems of plants and bacteria, respectively. The crystal structures of two members of this family have been determined: poplar leaf plastocyanin and Pseudomonas aeruginosa azurin. Both proteins contain two β-sheets, packed face-to-face. Using computed superpositions of the structures, we have aligned the sequences, identified homologous positions, and studied how the structures have changed as a result of mutations.The residues in the vicinity of the copper-binding site show minimal amino acid substitution and form almost identical structures. Other portions of these proteins are more variable in sequence and in structure. Buried residues tend to maintain their hydrophobic character, but mutations change their volume. The mean variation in volume of homologous buried residues is 54 Å3. The differences in size and shape of these buried residues are accommodated by a 3.8 Å shift in relative position of the packed β-sheets. This shift does not affect the copper binding site, because the residues that form this site are in, or adjacent to, just one of the β-sheets.  相似文献   

11.
The solid state secondary structure of myoglobin, RNase A, concanavalin A (Con A), poly(L -lysine), and two linear heterooligomeric peptides were examined by both far-uv CD spectroscopy1 and by ir spectroscopy. The proteins associated from water solution on glass and mica surfaces into noncrystalline, amorphous films, as judged by transmission electron microscopy of carbon-platinum replicas of surface and cross-fractured layer. The association into the solid state induced insignificant changes in the amide CD spectra of all α-helical myoglobin, decreased the molar ellipticity of the α/β RNase A, and increased the molar ellipticity of all-β Con A with no change in the positions of the bands' maxima. High-temperature exposure of the films induced permanent changes in the conformation of all proteins, resulting in less α-helix and more β-sheet structure. The results suggest that the protein α-helices are less stable in films and that the secondary structure may rearrange into β-sheets at high temperature. Two heterooligomeric peptides and poly (L -lysine), all in solution at neutral pH with “random coil” conformation, formed films with variable degrees of their secondary structure in β-sheets or β-turns. The result corresponded to the protein-derived Chou-Fasman amino acid propensities, and depended on both temperature and solvent used. The ir and CD spectra correlations of the peptides in the solid state indicate that the CD spectrum of a “random” structure in films differs from random coil in solution. Formic acid treatment transformed the secondary structure of the protein and peptide films into a stable α-helix or β-sheet conformations. The results indicate that the proteins aggregate into a noncrystalline, glass-like state with preserved secondary structure. The solid state secondary structure may undergo further irreversible transformations induced by heat or solvent. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
Analysis of electron-transfer (ET) kinetics data obtained from experiments on Ru-modified proteins (azurin, cytochrome c, myoglobin) and the bacterial photosynthetic reaction center reveals that distant donor-acceptor electronic couplings depend upon the secondary structure of the intervening polypeptide matrix. The β-sheet azurin structure efficiently and isotropically mediates coupling with an exponential distance-decay constant of 1.1?Å–1. The experimentally derived distance-decay constant of 1.4?Å–1 for long-range ET in myoglobin and the reaction center suggests that hydrogen-bond couplings are weaker through α helices than across β sheets. The donor-acceptor interactions of systems with comparable tunneling energies fall into two coupling zones: the β zone (bounded by distance-decay constants of 0.9?and 1.15 Å–1) includes all the β-sheet (azurin) couplings and all but one coupling in cytochrome c; the α zone (boundaries: 1.25 and 1.6?Å–1) includes less strongly coupled donor-acceptor pairs in myoglobin and the reaction center as well as a relatively weakly coupled pair in cytochrome c.  相似文献   

13.
Recent advances in three-dimensional electron microscopy (3D EM) have enabled the quantitative visualization of the structural building blocks of proteins at improved resolutions. We provide algorithms to detect the secondary structures (α-helices and β-sheets) from proteins for which the volumetric maps are reconstructed at 6–10 Å resolution. Additionally, we show that when the resolution is coarser than 10 Å, some of the supersecondary structures can be detected from 3D EM maps. For both these algorithms, we employ tools from computational geometry and differential topology, specifically the computation of stable/unstable manifolds of certain critical points of the distance function induced by the molecular surface. Our results connect mathematically well-defined constructions with bio-chemically induced structures observed in proteins.  相似文献   

14.
Abstract

We performed molecular dynamics simulations for various oligomers with different β-sheet conformations consisting of α-Synuclein 71–82 residues using an all atom force field and explicit water model. Tetramers of antiparallel β-sheet are shown to be stable, whereas parallel sheets are highly unstable due to the repulsive interactions between bulky and polar side chains as well as the weaker backbone hydrogen bonds. We also investigated the stabilities of double antiparallel β-sheets stacked with asymmetric and symmetric geometries. Our results show that this 12 amino acid residue peptide can form stable β-sheet conformers at 320K and higher temperatures. The backbone hydrogen bonds in β-sheet and the steric packing between hydrophobic side chains between β-sheets are shown to give conformational stabilities.  相似文献   

15.
Recently, it was suggested that parallel β-sheets have a significant dipole moment, in contrast to antiparallel sheets. Ab initio molecular-orbital (MO) calculations on parallel and antiparallel β-strands of tetra(Gly) show that they have very similar charge distributions. Interaction energies between two and three strands of tetra(Gly), obtained using the direct reaction field Hamiltonian, show that a particular choice of point charges is probably not crucial for calculating interactions within β-sheets, but that it might be for calculating interactions between these sheets and other parts of a protein, in particular, α-helices. The point-charge representation of our MO-SCF results will probably reduce the hazard of introducing artefacts in electrostatic calculations of protein conformational energies, provided the short-range interactions are treated in a more realistic way, i.e., such that intra- and interchain induction effects are included.  相似文献   

16.
The distance geometry approach for computing the tertiary structure of globular proteins emphasized in this series of papers (Goelet al., J. theor. Biol. 99, 705–757, 1982) is developed further. This development includes incorporation of some secondary structure information—the location of alpha helices in the primary sequence—in the algorithm to compute the tertiary structure of alpha helical globular proteins. An algorithm is developed which estimates the interresidue distances between chain-proximate helices. These distances, in conjunction with the global statistical average distances obtainable from a database of real proteins and determined by the primary sequence of the protein under study, are used to determine the tertiary structure. Five proteins, parvalbumin, hemerythrin, human hemoglobin, lamprey hemoglobin, and sperm whale myoglobin, are investigated. The root mean square (RMS) errors between the calculated structures and those determined by X-ray diffraction range from 4.78 to 7.56 Å. These RMSs are 0.21–2.76 Å lower than those estimated without the secondary structure information. Contact maps and three-dimensional backbone representations also show considerable improvements with the introduction of secondary structure information.  相似文献   

17.
We report the first direct observations of the physical and chemical microstructure of spider dragline, revealed by analytical transmission electron microscopy. Individual crystallites were imaged within the amorphous matrix. They are irregularly shaped, approximately 70–100 nm in diameter, and uniformly distributed throughout the matrix. Electron diffraction determined their space group to be P21. The corresponding orthogonal cell has lattice parameters of a = 13.31 Å (β-sheet repeat), b = 9.44 Å (interchain repeat within β-sheets), and c = 20.88 Å (repeat along polypeptide chain). Electron energy loss spectroscopy indicated compositional variations within the matrix, and between the crystallites and matrix. Most notably, calcium was found exclusively in the crystallites. Attempts to produce synthetic analogues of dragline, which exhibits an unparalleled combination of strength, stiffness, and toughness, cannot depend solely on duplicating the constituent proteins. The complex hierarchical microstructure of the natural material must be taken into account. © 1994 John Wiley & Sons, Inc.  相似文献   

18.
Robert W. Woody 《Biopolymers》1978,17(6):1451-1467
The rotational strength of the La transition in phenylalanine and tyrosine side chains has been calculated for dipeptides with various backbone and side-chain conformations. Similar calculations have also been performed for tripeptides in the β-turn conformation with aromatic residues at the corners of the turn. The interaction of the aromatic ring with neighboring peptides generates rotational strengths in the La transition of the order of 0.1 Debye-Bohr magneton. When the preferred backbone and side-chain conformations are considered, it is found that the most probable conformations have positive La bonds. This result accounts for the observation that the N-acyl amino acid amides of L -Tyr and L Phe have positive La bands. It also suggests that, although other interactions may affect the numerical value and even the sign, there will be a significant positive contribution to the rotational strength of aromatic residues in globular proteins from nearest-neighbor interactions. Calculations on proteins of known conformation at the nearest-neighbor level confirm the tendency toward positive La contributions for Phe and Tyr residues. This contribution can be of the order of 10% of the observed CD even in proteins with rather strong amide contributions. In some proteins, such as the gene 5 protein from bacteriophage fd and many snake-venom toxins, side-chain contributions from Tyr and Trp residues manifest themselves as positive CD bands in the 225–250-nm region. The magnitude of the nearest-neighbor contributions and the trend toward positive contributions are consistent with the observation of such CD bands in globular proteins. No special stacking interaction among aromatic side chains needs to be invoked.  相似文献   

19.
The structure of human lysozyme has been crystallographically refined at 1.5 Å resolution by difference map and restrained least-squares procedures to an R factor of 0.187. A comprehensive analysis of the non-bonded and hydrogen-bonded contacts in the lysozyme molecule, which were not restrained, revealed by the refinement has been carried out. The non-bonded CC contacts begin at ~3.45 Å, and the shorter contacts are dominated, as expected, by interactions between trigonal and tetrahedral carbon atoms. The CO contact distances have a “foot” at 3.05 Å. The CN distance plot shows a significant peak at 3.25 Å, which results from close contact between peptide NHs and carbonyl carbons involved in NiC′i ? 2 interactions in α-helices and reverse turns. The distances involving sulphur atoms discriminate SC trigonal interactions at 3.4 to 3.6 Å from SC tetrahedral interactions at 3.7 Å. All these types of non-bonded interactions show minimum distances close to standard van der Waals' separations.Analysis of hydrogen-bond distances has been carried out by using standard geometry to place hydrogen atoms and measuring the XHO distances. On this basis, there are 130 intramolecular hydrogens: 111 NHO bonds, of which 69 are between main-chain atoms, 13 between side-chain atoms and 29 between mainchain and side-chain atoms. If a cluster of four well-defined internal water molecules is included in the protein structure, there is a total of 19 OHO hydrogen bonds. The mean NO, NHO distances and HN?O angles are 2.96 ± 0.17 Å, 2.05 ± 0.18 Å and 18.5 ± 9.6 °, and the mean OO, OHO distances and HÔO angles are 2.83 ± 0.19 Å, 1.98 ± 0.26 Å and 23.8 ± 13.4 °. The distances agree well with standard values, although the hydrogen bonds are consistently more non-linear than in equivalent small molecules. An analysis of the hydrogen-bond angles at the receptor atom indicates that the α-helix, β-sheet and reverse turn have characteristic angular values. A detailed analysis of the regularity of the α-helices and reverse turns shows small but consistent differences between the α-helices in lysozyme and the current standard model, which may now need revision. Of the 21 reverse turns that include a hydrogen bond, the conformations of 19 agree very closely with four of the five standard types. We conclude that the restrained least-squares method of refinement has been validated by these analyses.  相似文献   

20.
The probability of occurrence of helix and β-sheet residues in 47 globular proteins was determined as a function of local hydrophobicity, which was defined by the sum of the Nozaki-Tanford transfer free energies at two nearest-neighbors on both sides of the amino acid sequence. In general, hydrophilic amino acids favor neither helix nor β-sheet formations when neighbor residues are also hydrophilic but favor helix formation at higher local hydrophobicity. On the other hand, some hydrophobic amino acids such as Met, Leu, and Ile favor helix formation when neighbor residues are hydrophilic. None of the hydrophobic amino acids favor β-sheet formation with hydrophilic neighbors, but most of them strongly favor β-sheet formation at high local hydrophobicity. When the average of 20 amino acids is taken, both helix and β-sheet residue probabilities are higher at higher local hydrophobicity, although the increase is steeper for β-sheets. Therefore, β-sheet formation is more influenced by local hydrophobicity than helix formation. Generally, helices are nearer the surface and tend to have hydrophilic and hydrophobic faces at opposite sides. The tendency of alternating regions of hydrophilic and hydrophobic residues in a helical sequence was revealed by calculating the correlation of the Nozaki-Tanford values. Such amphipathic helices may be important in protein–protein and protein–lipid interactions and in forming hydrophilic channels in the membrane. The choice of 30 nonhomologous proteins as the data set did not alter the above results.  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