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.
A matrix formulation of the conformational partition function has been used to examine helix ? sheet transitions in homopolyamino acids. α-Helices are weighted by Zimm-Bragg parameters σ and s. Antiparallel β-sheets with tight bends are weighted by the parameters t, δ, and τ, where t is the propagation parameter. In addition, each bend contributes a factor δ, and each residue in the sheet that does not have a partner in the preceding strand contributes a factor τ. The helix can be the dominant conformation in a long chain only if two conditions are satisfied simultaneously: (i) s > 1 , and (ii) either s > t, or σ, δ, and τ are assigned values that inflict a greater penalty on antiparallel sheets than on helices. The maximum amount of coil developed during the helix ? sheet transition is strongly influenced by the size of τ, but it is only weakly dependent on the size of δ. Previously reported optical rotatory dispersion, CD, laser Raman, and nmr studies of thermally induced α ? β transitions in homopolyamino acids, notably poly(L -lysine), demonstrate that little random coil is present. If the random coil content is to remain small during the helix ? sheet transition, τ must be significantly less than unity. A small value for τ means that there is a significant penalty assessed to lysyl residues in an antiparallel sheet that do not have a partner in a preceding strand.  相似文献   

2.
Hyun Joo  Jerry Tsai 《Proteins》2014,82(9):2128-2140
To understand the relationship between protein sequence and structure, this work extends the knob‐socket model in an investigation of β‐sheet packing. Over a comprehensive set of β‐sheet folds, the contacts between residues were used to identify packing cliques: sets of residues that all contact each other. These packing cliques were then classified based on size and contact order. From this analysis, the two types of four‐residue packing cliques necessary to describe β‐sheet packing were characterized. Both occur between two adjacent hydrogen bonded β‐strands. First, defining the secondary structure packing within β‐sheets, the combined socket or XY:HG pocket consists of four residues i, i+2 on one strand and j, j+2 on the other. Second, characterizing the tertiary packing between β‐sheets, the knob‐socket XY:H+B consists of a three‐residue XY:H socket (i, i+2 on one strand and j on the other) packed against a knob B residue (residue k distant in sequence). Depending on the packing depth of the knob B residue, two types of knob‐sockets are found: side‐chain and main‐chain sockets. The amino acid composition of the pockets and knob‐sockets reveal the sequence specificity of β‐sheet packing. For β‐sheet formation, the XY:HG pocket clearly shows sequence specificity of amino acids. For tertiary packing, the XY:H+B side‐chain and main‐chain sockets exhibit distinct amino acid preferences at each position. These relationships define an amino acid code for β‐sheet structure and provide an intuitive topological mapping of β‐sheet packing. Proteins 2014; 82:2128–2140. © 2014 Wiley Periodicals, Inc.  相似文献   

3.
A matrix treatment of the formation of intramolecular anti-parallel β-sheets from a statistical coil has been extended to incorporate interstrand loops of arbitrary size. The behavior of the model is compared with a simpler version in which all pairs of contiguous strands were connected by β-bends. When large interstrand loops are allowed, there are many more types of sheets than is the case when all contiguous strands must be connected by tight or β-bends. For this reason, the larger interstrand loops make it easier to introduce the initial sheet into a statistical coil, and the sheet content is enhanced in the early stages of stages of sheet formation (i.e., at small values of the growth parameter t). As the transition continues (i.e., as t increases), a stage will be reached where occupancy of the statistical coil state is negligible because nearly all residues are in sheets or interstrand loops. Now, additional sheet formation can be accomplished only at the expense of residues in the interior of interstrand loops. For this reason, the larger interstrand loops make it more difficult to complete the final stages of sheet formation. These effects are especially dramatic in the formation of cross-β-sheets.  相似文献   

4.
The formation of amyloid fibers and their deposition in the body is a characteristic of a number of devastating human diseases. Here, we propose a structural model, based on X-ray diffraction data, for the basic structure of an amyloid fibril formed by using the variants of the B1 domain of IgG binding protein G of Streptococcus. The model for the fibril incorporates four beta sheets in a bundle with a diameter of 45 A. Its cross-section, or layer, consists of four strands, one strand from each sheet. Layers stack on top of each other to form the fibril, which has an overall helical twist with a periodicity of about 154 A. Each strand interacts in a parallel fashion with the strands in the layers above and below it, in an infinite beta sheet. Some geometric features of this model and the logic behind it may be applicable for constructing other related cross-beta amyloid fibrils.  相似文献   

5.
We have systematically mutated residues located in turns between beta-strands of the intestinal fatty acid binding protein (IFABP), and a glycine in a half turn, to valine and have examined the stability, refolding rate constants and ligand dissociation constants for each mutant protein. IFABP is an almost all beta-sheet protein exhibiting a topology comprised of two five-stranded sheets surrounding a large cavity into which the fatty acid ligand binds. A glycine residue is located in seven of the eight turns between the antiparallel beta-strands and another in a half turn of a strand connecting the front and back sheets. Mutations in any of the three turns connecting the last four C-terminal strands slow the folding and decrease stability with the mutation between the last two strands slowing folding dramatically. These data suggest that interactions between the last four C-terminal strands are highly cooperative, perhaps triggered by an initial hydrophobic collapse. We suggest that this trigger is collapse of the highly hydrophobic cluster of amino acids in the D and E strands, a region previously shown to also affect the last stage of the folding process (Kim et al., 1997). Changing the glycine in the strand between the front and back sheets also results in a unstable, slow folding protein perhaps disrupting the D-E strand interactions. For most of the other turn mutations there was no apparent correlation between stability and refolding rate constants. In some turns, the interaction between strands, rather than the turn type, appears to be critical for folding while in others, turn formation itself appears to be a rate limiting step. Although there is no simple correlation between turn formation and folding kinetics, we propose that turn scanning by mutagenesis will be a useful tool for issues related to protein folding.  相似文献   

6.
Koch O  Bocola M  Klebe G 《Proteins》2005,61(2):310-317
A systematic analysis of the hydrogen-bonding geometry in helices and beta sheets has been performed. The distances and angles between the backbone carbonyl O and amide N atoms were correlated considering more than 1500 protein chains in crystal structures determined to a resolution better than 1.5 A. They reveal statistically significant trends in the H-bond geometry across the different secondary structural elements. The analysis has been performed using Secbase, a modular extension of Relibase (Receptor Ligand Database) which integrates information about secondary structural elements assigned to individual protein structures with the various search facilities implemented into Relibase. A comparison of the mean hydrogen-bond distances in alpha helices and 3(10) helices of increasing length shows opposing trends. Whereas in alpha helices the mean H-bond distance shrinks with increasing helix length and turn number, the corresponding mean dimension in 3(10) helices expands in a comparable series. Comparing similarly the hydrogen-bond lengths in beta sheets there is no difference to be found between the mean H-bond length in antiparallel and parallel beta sheets along the strand direction. In contrast, an interesting systematic trend appears to be given for the hydrogen bonds perpendicular to the strands bridging across an extended sheet. With increasing number of accumulated strands, which results in a growing number of back-to-back piling hydrogen bonds across the strands, a slight decrease of the mean H-bond distance is apparent in parallel beta sheets whereas such trends are obviously not given in antiparallel beta sheets. This observation suggests that cooperative effects mutually polarizing spatially well-aligned hydrogen bonds are present either in alpha helices and parallel beta sheets whereas such influences seem to be lacking in 3(10) helices and antiparallel beta sheets.  相似文献   

7.
Theoretical analysis was carried out to determine how the approximately 20% of beta-structure observed in the 18.5 kilodalton (kDa) myelin basic protein (MBP) could be organized into a relatively stable beta-sheet. The beta-sheet is presumed to consist of the five most hydrophobic segments of polypeptide chain, which have beta-structure potential. These correspond approximately to sequences 15-21, 37-45, 84-92, 106-112, and 148-154 (rabbit MBP sequence numbering) and constitute beta-strands a, b, c, d, and e, respectively. A number of constraints are imposed upon the sheet; e.g., it should have the same topology in all MBP forms (21.5, 18.5, 17, and 14 kDa); strand e should lie at the sheet edge; strands b, c, and d should be ordered sequentially; the sheet formed by strands a, b, c, and d should be antiparallel; a maximum of the nonpolar surface area should be removed from the aqueous milieu; and charged side chains should be solvent-accessible. On the basis of these constraints it is possible to propose six orthogonally packed beta-sheets having different topologies. If strand e is restricted to an antiparallel alignment, the number of different sheets is reduced to four. Each of these sheets can form a relatively compact hydrophobic globular region. Two of the strands (a and e) can undergo transitions to alpha-helix without disrupting the structure of the remaining sheet bcd or producing major topologic rearrangements of the polypeptide chain.  相似文献   

8.
A new model structure is proposed for the silk I form of the crystalline domains of Bombyx mori silk fibroin and the corresponding crystal form of poly(L-Ala-Gly). It was deduced from conformational energy computations on stacked sheet structures of poly(L-Ala-Gly). The novel sheet structure contains interstrand hydrogen bonds but is composed of anti-parallel polypeptide chains whose conformation differs from that of the antiparallel beta-sheets that constitute the silk II structure. The strands of the new sheet have a two-residue repeat, in which the Ala residues adopt a right-handed and the Gly residues a left-handed sheet-like conformation. The computed unit cell is orthorhombic, with cell dimensions a = 8.94 A, b = 6.46 A, and c = 11.26 A. The model accounts for most spacings in the observed fiber x-ray diffraction patterns of silk I and of the silk-I-like form of poly(L-Ala-Gly), and it is consistent with nmr and ir spectroscopic data. As a test of the computations, the well-established beta-sheet structure of silk II and the corresponding form of poly(L-Ala-Gly) have been reproduced. The computed energies for the two forms of poly(L-Ala-Gly) indicate that the silk-II-like form is more stable, by about 1.0 kcal/mol per residue. The main difference between the two structures is the orientation of the Ala side chains of neighboring strands in each sheet. In the Pauling-Corey beta-sheet and in the silk II form, referred to as an "in-register" structure, the Ala side chains of every strand point to the same side of a sheet. In the silk I structure, referred to as "out-of-register," the side chains of Ala residues in adjacent strands point to opposite sides of the sheet.  相似文献   

9.
A program that generates amino acid sequences that are compatiblewith the Greek key protein motif is presented. Using statisticaldata derived from the structures of molecules from the proteindatabank, the novel algorithm generates amino acid sequencescompatible with an 8-stranded perfect Greek key jellyroll motifIn this motif, all hydrogen bonds present in the theoreticaloriginating (ß-hairpin stay in register as the whole8-stranded domain fokis at once in an anti clockwise swirl.Eight residues are generated per strand and 32 residues persheet making 64 residues in the antiparallel (ß-barrel.The seven loops between ß-stands contain an additional27 residues. All recognized features of (ß-sheetsand ß-strands such as alternating hydrophobic, hydrophilicresidues with hydrophobics on the narrow-hydrogen-bond-pair,concave side of the theoretical originating (ß-ribbon;sheet twist, strand twist, side chain rotation about the strands;the theories of side chain packing between the sheets; an average30° rotation between (ß-sheets; the theoreticalanti complementary patch residues of each sheet; and the anticomplementaty,isotropically stressed hyperbolic parabloid shape of each sheet,are taken into account in the program. The sequences of theloops between strands are designed by turn type and strand twistis considered in the design of the motif's single (ß-hairpinturn. Secondary structure parameters and between-strand aminoacid pair correlations also figure importantly in the novelalgorithm.  相似文献   

10.
BACKGROUND: Both backbone hydrogen bonding and interactions between sidechains stabilize beta sheets. Cross-strand interactions are the closest contacts between the sidechains of a beta sheet. Here we investigate the energetics of cross-strand interactions using a variant of the B1 domain of immunoglobulin G (IgG) binding protein G (beta1) as our model system. RESULTS: Pairwise mutations of polar and nonpolar residues were made at a solvent-exposed site between the two central parallel beta strands of beta1. Both stabilizing and destabilizing interactions were measured. The greatest stabilizations were observed for charge-charge interactions. Our experimental study of sidechain interactions correlates with statistical preferences: residue pairs for which we measure stabilizing interaction energies occur together frequently, whereas destabilizing pairs are rarely observed together. CONCLUSIONS: Sidechain interactions modulate the stability of beta sheets. We propose that cross-strand sidechain interactions specify correct strand register and ordering through the energetic benefit of optimally arranged pairings.  相似文献   

11.
The structure of the gene 5 DNA unwinding protein from bacteriophage fd has been determined by X-ray diffraction analysis of single crystals to 2.3 Å resolution using six isomorphous heavy-atom derivatives. The essentially globular monomer appears to consist of three secondary structural elements, a radically twisted three-stranded antiparallel β sheet and two distinct anti-parallel β loops, which are joined by short segments of extended polypeptide chain. The molecule contains no α-helix. A long groove, or arch, 30 Å in length is formed by the underside of the twisted β sheet and one of the two β ribbons. We believe this groove to be the DNA binding region, and this is supported by the assignment of residues on its surface implicated in binding by solution studies. These residues include several aromatic amino acids which may intercalate or stack upon the bases of the DNA. Two monomers are maintained as a dimer by the very close interaction of symmetry related β ribbons about the molecular dyad. About six residues at the amino and carboxyl terminus are in extended conformation and both seem to exhibit some degree of disorder. The amimo-terminal methionine is the locus for binding the platinum heavy-atom derivatives and tyrosine 26 for attachment of the major iodine substituent.  相似文献   

12.
Crystal structure of cat muscle pyruvate kinase at a resolution of 2.6 A   总被引:23,自引:0,他引:23  
The structure of pyruvate kinase (EC 2.7.1.40) has been determined from a 2.6 Å resolution electron density map. This map shows more detail than the previous 3.1 Å map (Stammers &; Muirhead, 1977) and has enabled a detailed chain folding to be established for two out of the three domains which make up each of the four identical subunits. A provisional chain folding has been established for the third domain. The results have been briefly reported in a previous paper (Levine et al., 1978). Details of the structure determination and a further discussion of the results are presented in this paper.Domain A (the three domains of pyruvate kinase are referred to as A, B and C) can be described in terms of a cylindrical eight-stranded parallel β sheet and an outer coaxial cylinder of eight α helices. The α helices connect adjacent strands of the β sheet. Domain B is made up of a closed anti-parallel β sheet structure. Domain C is a five-stranded β sheet of which the fourth strand is anti-parallel and the rest parallel. These strands are also interconnected by α helices.Domain A can be dissected into eight consecutive β strand—α helix units starting from the N-terminus. The arrangement of these relative to each other can be most simply described by relating them to eight planes, each at 40 ° to the cylinder axis and symmetrically placed around the cylinder. When unit 2 is aligned with one of these planes then units 1, 3, 4, 5 and 8 are also closely aligned with a plane. This analysis is also applied to triosephosphate isomerase and a strikingly similar arrangement is found. A detailed comparison of the two structures is presented. Although the lack of a chemical sequence makes it difficult to identify the amino acid residues of pyruvate kinase, side-chains are clearly visible in the map and this information is correlated with the results of previous 6 Å substrate soaking experiments and with the structure of triosephosphate isomerase. The similarities and differences are discussed in terms of similarities and differences in the reactions catalysed and also of different subunit packing.  相似文献   

13.
We present an algorithm that is able to propose compact models of protein 3D structures, only starting from the prediction of the nature and length of regular secondary structures. Helices are modeled by cylinders and sheets by helicoid surfaces, all strands of a sheet being considered as a single block. It means that relative topology of the strands inside one sheet is a prerequisite. Loops are only considered as constraints, given by the maximal distance between their Calpha extremities according to their sequence length. Unconnected regular secondary structures are reduced to a single point, the center of their hydrophobic faces. These centers are then repeatedly moved in order to obtain a compact hydrophobic core. To prevent secondary structures from interpenetrating, a repulsive term is introduced in the function whose minimization leads to the compact structure. This RUSSIA (Rigid Unconnected Secondary Structure Assembly) algorithm has the advantage of relying on a small number of variables and therefore many initial conformations can be tested. Flexibility is produced in the following way: helices or sheets are allowed to rotate around the direction leading to the center of the model; residues in a sheet can slide along the main direction of the strand where they are embedded. RUSSIA is fast and simple and it produces on a test set several neighbor good models with an r.m.s. to the native structures in the range 1.4-3.7 A. These models can be further treated by statistical potentials used in threading approaches in order to detect the best candidate. The limits of the present method are the following: small proteins with few secondary structures are excluded; multi domain proteins must be split into several compact globular domains from their sequences; sheets of more than five strands and completely buried helices are not treated. In this first paper the algorithm is developed and in Part II, which follows, some applications are presented and the program is evaluated.  相似文献   

14.
Bacterial cell wall peptidoglycans are built from unbranched β-(1 → 4)-linked glycan chains composed of alternately repeating units of N-acetylglucosamine and N-acetylmuramic acid residues, with peptide side chains attached to the muramic acid residues. The glycan chains are interconnected by peptide bonds formed between the peptide side chains. Through the use of three-dimensional molecular models, two configurations of the glycan strands and the peptide side chains are described, which by their constancy of form reflect the fundamental constancies of the covalent structures. Each of these two models will accommodate any chemical modification that has been observed in bacteria without change in the configuration of the peptide backbone. Some alterations in the chemical structure, which have been sought in bacteria, but not found, would not be tolerated by the models. In these models, glycan strands are parallel, with their lengths and widths predominantly in the plane of the cell wall. The cross-bridging portions of the peptide side chains are at right angles to the glycan strand, in a separate, parallel plane. A compact model is presented in which the peptide side chain is closely appressed to the glycan strand and is stabilized by three hydrogen bonds per disaccharide–peptide subunit. In a second model, the peptide side chain is raised away from the glycan strand in an entirely extended configuration. The compact and extended forms are interconvertible. The thickness of a sheet of peptidoglycan would be from 10.6 to 11.1 Å for the compact model, and 19.1 Å for the extended model.  相似文献   

15.
We analysed the length distributions of different types of beta-strand in a high resolution, non-homologous set of 500 protein structures, finding differences in their mean lengths. Antiparallel edge strands in strand-turn-strand motifs show a preference for an even number of residues. This propensity is enhanced if the length is corrected for beta-bulges, which insert an extra residue into the strand. Residues in antiparallel edge beta-strands alternate between being in hydrogen bonded and non-hydrogen bonded rings. Antiparallel edges with an even number of residues are more likely to have their final beta residue in a non-hydrogen bonded ring. This suggests that non-hydrogen bonded rings are intrinsically more stable than hydrogen bonded rings, perhaps because its side chain packing is closer. Therefore, we suggest that a simple way to increase beta-hairpin stability, or the stability of an antiparallel edge strand, is to have a non-hydrogen bonded ring at the end of the strand.  相似文献   

16.
17.
The molecular structure of endothiapepsin (EC 3.4.23.6), the aspartic proteinase from Endothia parasitica, has been refined to a crystallographic R-factor of 0.178 at 2.1 A resolution. The positions of 2389 protein non-hydrogen atoms have been determined and the present model contains 333 solvent molecules. The structure is bilobal, consisting of two predominantly beta-sheet domains that are related by an approximate 2-fold axis. Of approximately 170 residues, 65 are topologically equivalent when one lobe is superimposed on the other. Twenty beta-strands are arranged as five beta-sheets and are connected by regions involving 29 turns and four helices. A central sheet involves three antiparallel strands from each lobe organized around the dyad axis. Each lobe contains a further local dyad that passes through two sheets arranged as a sandwich and relates two equivalent motifs of four antiparallel strands (a, b, c, d) followed by a helix or an irregular helical region. Sheets 1N and 1C, each contain two interpenetrating psi structures contributed by strands c,d,d' and c',d',d, which are related by the intralobe dyad. A further sheet, 2N or 2C, is formed from two extended beta-hairpins from strands b,c and b',c' that fold above the sheets 1N and 1C, respectively, and are hydrogen-bonded around the local intralobe dyad. Asp32 and Asp215 are related by the interlobe dyad and form an intricate hydrogen-bonded network with the neighbouring residues and comprise the most symmetrical part of the structure. The side-chains of the active site aspartate residues are held coplanar and the nearby main chain makes a "fireman's grip" hydrogen-bonding network. Residues 74 to 83 from strands a'N and b'N in the N-terminal lobe form a beta-hairpin loop with high thermal parameters. This "flap" projects over the active site cleft and shields the active site from the solvent region. Shells of water molecules are found on the surface of the protein molecule and large solvent channels are observed within the crystal. There are only three regions of intermolecular contacts and the crystal packing is stabilized by many solvent molecules forming a network of hydrogen bonds. The three-dimensional structure of endothiapepsin is found to be similar to two other fungal aspartic proteinases, penicillopepsin and rhizopuspepsin. Even though sequence identities of endothiapepsin with rhizopuspepsin and penicillopepsin are only 41% and 51%, respectively, a superposition of the three-dimensional structures of these three enzymes shows that 237 residues (72%) are within a root-mean-square distance of 1.0 A.  相似文献   

18.
Two separate unrefined models for the secondary structure of two subfamilies of the 6-phospho-β-D -galactosidase superfamily were independently constructed by examining patterns of variation and conservation within homologous protein sequences, assigning surface, interior, parsing, and active site residues to positions in the alignment, and identifying periodicities in these. A consensus model for the secondary structure of the entire superfamily was then built. The prediction tests the limits of an unrefined prediction made using this approach in a large protein with substantial functional and sequence divergence within the family. The protein belongs to the (α–β class), with the core β strands aligned parallel. The supersecondary structural elements that are readily identified in this model is a parallel β sheet built by strands C, D, and E, with helices 2 and 3 connecting strands (C + D) and (D + E), respectively, and an analogous α–β unit (strand G and helix 7) toward the end of the sequence. The resemblance of the supersecondary model to the tertiary structure formed by 8-fold α–β barrel proteins is almost certainly not coincidental. © 1995 Wiley-Liss, Inc.  相似文献   

19.
The crystallographic investigation of the retro-inverso peptide Bz-S-gAla-R-mAla-NHPh reveals an extended backbone conformation where the NH groups of the gem-diamino alkyl moiety and the CO groups of the malonyl residue face side by side. This extended conformation, presenting all carbonyls on opposite sides of the NH groups, is stabilized by interstrand H-bonds running in a single direction of the parallel beta-sheets that characterize the crystal packing. These sheets differ from the beta-sheets formed by native amino acids only. (1)H-NMR nuclear Overhauser effect spectroscopy (NOESY) experiments suggest that a conformation similar to that found in the crystal also prevails in dimethylsulfoxide solution. Previous potential energy calculations of gem-diamino alkyl (g) and malonyl (m) Ala residues predicted that extended forms were less stable than the helical ones because of strong electrostatic repulsions between the parallel polar groups. Similar arguments were invoked to give more weight to helical forms of the retro-peptide units in the proposal of packing models of some nylons in their crystalline polar regions. The present findings show that both g and m Ala residues can experience the extended conformation in the beta-sheet aggregation. The energy increase occurring in one strand, due to the parallel orientation of consecutive peptide dipoles, is more than compensated by favorable cooperative interactions among head-to-tail aligned peptide dipoles of facing strands, resulting in the formation of two C==O...H==N H-bonds per residue.  相似文献   

20.
A systematic survey of seven parallel alpha/beta barrel protein domains, based on exhaustive structural comparisons, reveals that a sizable proportion of the alpha beta loops in these proteins--20 out of a total of 49--belong to either one of two loop types previously described by Thornton and co-workers. Six loops are of the alpha beta 1 type, with one residue between the alpha-helix and beta-strand, and 13 are of the alpha beta 3 type, with three residues between the helix and the strand. Protein fragments embedding the identified loops, and termed alpha beta connections since they contain parts of the flanking helix and strand, have been analyzed in detail revealing that each type of connection has a distinct set of conserved structural features. The orientation of the beta-strand relative to the helix and loop portions is different owing to a very localized difference in backbone conformation. In alpha beta 1 connections, the chain enters the beta-strand via a residue adopting an extended conformation, while in alpha beta 3 it does so via a residue in a near alpha-helical conformation. Other conserved structural features include distinct patterns of side chain orientation relative to the beta-sheet surface and of main chain H-bonds in the loop and the beta-strand moieties. Significant differences also occur in packing interactions of conserved hydrophobic residues situated in the last turn of the helix. Yet the alpha-helix surface of both types of connections adopts similar orientations relative to the barrel sheet surface. Our results suggest furthermore that conserved hydrophobic residues along the sequence of the connections, may be correlated more with specific patterns of interactions made with neighboring helices and sheet strands than with helix/strand packing within the connection itself. A number of intriguing observations are also made on the distribution of the identified alpha beta 1 and alpha beta 3 loops within the alpha/beta-barrel motifs. They often occur adjacent to each other; alpha beta 3 loops invariably involve even numbered beta-strands, while alpha beta 1 loops involve preferentially odd beta-strands; all the analyzed proteins contain at least one alpha beta 3 loop in the first half of the eightfold alpha/beta barrel. Possible origins of all these observations, and their relevance to the stability and folding of parallel alpha/beta barrel motifs are discussed.  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