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.
Folding type-specific secondary structure propensities of 20 naturally occurring amino acids have been derived from α-helical, β-sheet, α/β, and α+β proteins of known structures. These data show that each residue type of amino acids has intrinsic propensities in different regions of secondary structures for different folding types of proteins. Each of the folding types shows markedly different rank ordering, indicating folding type-specific effects on the secondary structure propensities of amino acids. Rigorous statistical tests have been made to validate the folding type-specific effects. It should be noted that α and β proteins have relatively small α-helices and β-strands forming propensities respectively compared with those of α+β and α/β proteins. This may suggest that, with more complex architectures than α and β proteins, α+β and α/β proteins require larger propensities to distinguish from interacting α-helices and β-strands. Our finding of folding type-specific secondary structure propensities suggests that sequence space accessible to each folding type may have differing features. Differing sequence space features might be constrained by topological requirement for each of the folding types. Almost all strong β-sheet forming residues are hydrophobic in character regardless of folding types, thus suggesting the hydrophobicities of side chains as a key determinant of β-sheet structures. In contrast, conformational entropy of side chains is a major determinant of the helical propensities of amino acids, although other interactions such as hydrophobicities and charged interactions cannot be neglected. These results will be helpful to protein design, class-based secondary structure prediction, and protein folding. © 1998 John Wiley & Sons, Inc. Biopoly 45: 35–49, 1998  相似文献   

2.
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   

3.
Heterodimeric integrin adhesion receptors regulate cell migration, survival and differentiation in metazoa by communicating signals bi‐directionally across the plasma membrane. Protein engineering and mutagenesis studies have suggested that the dissociation of a complex formed by the single‐pass transmembrane (TM) segments of the α and β subunits is central to these signalling events. Here, we report the structure of the integrin αIIbβ3 TM complex, structure‐based site‐directed mutagenesis and lipid embedding estimates to reveal the structural event that underlies the transition from associated to dissociated states, that is, TM signalling. The complex is stabilized by glycine‐packing mediated TM helix crossing within the extracellular membrane leaflet, and by unique hydrophobic and electrostatic bridges in the intracellular leaflet that mediate an unusual, asymmetric association of the 24‐ and 29‐residue αIIb and β3 TM helices. The structurally unique, highly conserved integrin αIIbβ3 TM complex rationalizes bi‐directional signalling and represents the first structure of a heterodimeric TM receptor complex.  相似文献   

4.
The maturation of connective tissue involves the organization of collagen fibres by resident fibroblasts. Fibroblast attachment to collagen has been demonstrated to involve cell surface receptors, integrins of the β1 family. Integrins are associated with cytoplasmic actin of microfilaments either directly or through focal adhesions. The major actin isoform of fibroblast microfilaments is β actin and to a lesser extent α smooth muscle (α SM) actin. Cultured human dermal fibroblasts derived from adult dermis, newborn foreskin or keloid scar were grown on either uncoated or collagen-coated surfaces. The expression and synthesis of both α2β1 integrin and α SM actin were followed by immunohistology and immunoprecipitation. Fibroblasts on uncoated surfaces expressed little α2β1 integrin on their surface, while 20 per cent of them demonstrated α SM actin within microfilaments. Fibroblasts grown on a collagen-coated surface minimally expressed α SM actin in microfilament structures and a majority of the cells were positive for α2β1 integrin on their membranes. Using [35S]-methionine incorporation and immunoprecipitation, it was shown that fibroblasts grown in uncoated dishes synthesized more α SM actin than fibroblasts grown on collagen-coated dishes. In contrast, fibroblasts grown on collagen coated dishes synthesized more α2β1 integrin compared to the same cells grown on uncoated dishes. Fibroblasts maintained on a type I collagen upregulate the expression and synthesis of α2β1 integrin, and downregulate the expression and synthesis of α SM actin. © 1998 John Wiley & Sons, Ltd.  相似文献   

5.
The conformational preferences of the monomeric units of a series of poly(α-alkyl-β-L-aspartate)s were examined by quantum mechanical calculations. α-Alkyl-β-aspartamyl m-oligopeptides with a number of residues m ranging from 1 to 7 and arranged in the conformations experimentally observed for their corresponding polymers were computed. Both their total relative energies and their cooperative energy differences were compared as a function of the length of the oligopeptide and the nature of the alkyl side group. Results revealed that the 13/4 helical arrangement is the most stable structure for the isolated polymer chain for any side group, although a 17/4 helix becomes favored in the case of methyl and ethyl groups due to the packing effects. On the other hand, the stability of the 4/1 helix appears to be the preferred conformation for side groups with a branched constitution. © 1997 John Wiley & Sons, Inc. Biopoly 41: 721–729, 1997  相似文献   

6.
In vivo regeneration of lost or dysfunctional islet β cells can fulfill the promise of improved therapy for diabetic patients. To achieve this, many mitogenic factors have been attempted, including gamma‐aminobutyric acid (GABA). GABA remarkably affects pancreatic islet cells’ (α cells and β cells) function through paracrine and/or autocrine binding to its membrane receptors on these cells. GABA has also been studied for promoting the transformation of α cells to β cells. Nonetheless, the gimmickry of GABA‐induced α‐cell transformation to β cells has two different perspectives. On the one hand, GABA was found to induce α‐cell transformation to β cells in vivo and insulin‐secreting β‐like cells in vitro. On the other hand, GABA treatment showed that it has no α‐ to β‐cell transformation response. Here, we will summarize the physiological effects of GABA on pancreatic islet β cells with an emphasis on its regenerative effects for transdifferentiation of islet α cells to β cells. We will also critically discuss the controversial results about GABA‐mediated transdifferentiation of α cells to β cells.  相似文献   

7.
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   

8.
9.
The acid-unfolded state of equine β-lactoglobulin was characterized by means of circular dichroism, nuclear magnetic resonance, analytical gel-filtration chromatography, and analytical centrifugation. The acid-unfolded state of equine β-lactoglobulin has a substantial secondary structure as shown by the far-ultraviolet circular dichroism spectrum but lacks persistent tertiary packing of the side chains as indicated by the near-ultraviolet circular dichroism and nuclear magnetic resonance spectra. It is nearly as compact as the native conformation as shown by the gel filtration and sedimentation experiments, and it has the exposed hydrophobic surface as indicated by its tendency to aggregate. All of these characteristics indicate that the acid-unfolded state of equine β-lactoglobulin is a molten globule state. The α helix content in the acid-unfolded state, which has been estimated from the circular dichroism spectrum, is larger than that in the native state, suggesting the presence of nonnative α helices in the molten globule state. This result suggests the generality of the intermediate with nonnative α helices during the folding of proteins having the β-clam fold. © 1997 Wiley-Liss Inc.  相似文献   

10.
The three-dimensional structure of the Sorghum bicolor seed protein γ-thionin SIα1 has been determined by 2D 1H nuclear magnetic resonance (NMR) spectroscopy. The secondary structure of this 47-residue antifungal protein with four disulphide bridges consists of a three-stranded antiparallel sheet and one helix. The helix is tethered to the sheet by two disulphide bridges which link two successive turns of the helix to alternate residues i, i + 2 in one strand. Possible binding sites for antifungal activity are discussed. The same fold has been observed previously in several scorpion toxins. Proteins 32:334–349, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

11.
Examination of crystal structures of restriction endonucleases EcoRI and EcoRV complexes with their cognate DNA revealed a common structural element, which forms the core of both proteins. This element consists of a five-stranded β-sheet and two α-helices packed against it and could be described as α–β sandwich in which helices and β-strands lie in two stacked layers. While the spatial structure of this α–β sandwich is conserved in both enzymes, there are no detectable similarities between amino acid sequences except of a few residues involved in active site formation. Probably, other restriction endonucleases which have similar organization of the active site might possess similar structural element regardless of DNA sequence recognized and recognition elements in the enzyme used. © 1994 Wiley-Liss, Inc.  相似文献   

12.
Integral membrane proteins (of the α-helical class) are of central importance in a wide variety of vital cellular functions. Despite considerable effort on methods to predict the location of the helices, little attention has been directed toward developing an automatic method to pack the helices together. In principle, the prediction of membrane proteins should be easier than the prediction of globular proteins: there is only one type of secondary structure and all helices pack with a common alignment across the membrane. This allows all possible structures to be represented on a simple lattice and exhaustively enumerated. Prediction success lies not in generating many possible folds but in recognizing which corresponds to the native. Our evaluation of each fold is based on how well the exposed surface predicted from a multiple sequence alignment fits its allocated position. Just as exposure to solvent in globular proteins can be predicted from sequence variation, so exposure to lipid can be recognized by variable-hydrophobic (variphobic) positions. Application to both bacteriorhodopsin and the eukaryotic rhodopsin/opsin families revealed that the angular size of the lipid-exposed faces must be predicted accurately to allow selection of the correct fold. With the inherent uncertainties in helix prediction and parameter choice, this accuracy could not be guaranteed but the correct fold was typically found in the top six candidates. Our method provides the first completely automatic method that can proceed from a scan of the protein sequence databanks to a predicted three-dimensional structure with no intervention required from the investigator. Within the limited domain of the seven helix bundle proteins, a good chance can be given of selecting the correct structure. However, the limited number of sequences available with a corresponding known structure makes further characterization of the method difficult. © 1994 John Wiley & Sons, Inc.  相似文献   

13.
A good model to experimentally explore evolutionary hypothesis related to enzyme function is the ancient‐like dual‐substrate (βα)8 phosphoribosyl isomerase A (PriA), which takes part in both histidine and tryptophan biosynthesis in Streptomyces coelicolor and related organisms. In this study, we determined the Michaelis–Menten enzyme kinetics for both isomerase activities in wild‐type PriA from S. coelicolor and in selected single‐residue monofunctional mutants, identified after Escherichia coli in vivo complementation experiments. Structural and functional analyses of a hitherto unnoticed residue contained on the functionally important β → α loop 5, namely, Arg139, which was postulated on structural grounds to be important for the dual‐substrate specificity of PriA, is presented for the first time. Indeed, enzyme kinetics analyses done on the mutant variants PriA_Ser81Thr and PriA_Arg139Asn showed that these residues, which are contained on β → α loops and in close proximity to the N‐terminal phosphate‐binding site, are essential solely for the phosphoribosyl anthranilate isomerase activity of PriA. Moreover, analysis of the X‐ray crystallographic structure of PriA_Arg139Asn elucidated at 1.95 Å herein strongly implicates the occurrence of conformational changes in this β → α loop as a major structural feature related to the evolution of the dual‐substrate specificity of PriA. It is suggested that PriA has evolved by tuning a fine energetic balance that allows the sufficient degree of structural flexibility needed for accommodating two topologically dissimilar substrates—within a bifunctional and thus highly constrained active site—without compromising its structural stability.  相似文献   

14.
To investigate the role of α helices in protein thermostability, we compared energy characteristics of α helices from thermophilic and mesophilic proteins belonging to four protein families of known three-dimensional structure, for at least one member of each family. The changes in intrinsic free energy of α-helix formation were estimated using the statistical mechanical theory for describing helix/coil transitions in peptide helices [Munoz, V., Serrano, L. Nature Struc. Biol. 1:399–409, 1994; Munoz, V., Serrano, L. J. Mol. Biol. 245:275–296, 1995; Munoz, V., Serrano, L. J. Mol. Biol. 245:297–308, 1995]. Based on known sequences of mesophilic and thermophilic RecA proteins we found that (1) a high stability of α helices is necessary but is not a sufficient condition for thermostability of RecA proteins, (2) the total helix stability, rather than that of individual helices, is the factor determining protein thermostability, and (3) two facets of intrahelical interactions, the intrinsic helical propensities of amino acids and the side chain–side chain interactions, are the main contributors to protein thermostability. Similar analysis applied to families of L-lactate dehydrogenases, seryl-tRNA synthetases, and aspartate amino transferases led us to conclude that an enhanced total stability of α helices is a general feature of many thermophilic proteins. The magnitude of the observed decrease in intrinsic free energy on α-helix formation of several thermoresistant proteins was found to be sufficient to explain the experimentally determined increase of their thermostability. Free energies of intrahelical interactions of different RecA proteins calculated at three temperatures that are thought to be close to its normal environmental conditions were found to be approximately equal. This indicates that certain flexibility of RecA protein structure is an essential factor for protein function. All RecA proteins analyzed fell into three temperature-dependent classes of similar α-helix stability (ΔGint = 45.0 ± 2.0 kcal/mol). These classes were consistent with the natural origin of the proteins. Based on the sequences of protein α helices with optimized arrangement of stabilizing interactions, a natural reserve of RecA protein thermoresistance was estimated to be sufficient for conformational stability of the protein at nearly 200°C. Proteins 29:309–320, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

15.
An analysis of the amino acid distributions at 15 positions, viz., N“, N′, Ncap, N1, N2, N3, N4, Mid, C4, C3, C2, C1, Ccap, C′, and C” in 1,131 α-helices reveals that each position has its own unique characteristics. In general, natural helix sequences optimize by identifying the residues to be avoided at a given position and minimizing the occurrence of these avoided residues rather than by maximizing the preferred residues at various positions. Ncap is most selective in its choice of residues, with six amino acids (S, D, T, N, G, and P) being preferred at this position and another 11 (V, I, F, A, K, L, Y, R, E, M, and Q) being strongly avoided. Ser, Asp, and Thr are all more preferred at Ncap position than Asn, whose role at helix N-terminus has been highlighted by earlier analyses. Furthermore, Asn is also found to be almost equally preferred at helix C-terminus and a novel structural motif is identified, involving a hydrogen bond formed by Nδ2 of Asn at Ccap or C1 position, with the backbone carbonyl oxygen four residues inside the helix. His also forms a similar motif at the C-terminus. Pro is the most avoided residue in the main body (N4 to C4 positions) and at C-ter-minus, including Ccap of an α-helix. In 1,131 α-helices, no helix contains Pro at C3 or C2 positions. However, Pro is highly favoured at N1 and C′. The doublet X-Pro, with Pro at C′ position and extended backbone conformation for the X residue at Ccap, appears to be a common structural motif for termination of α-helices, in addition to the Schellman motif. Main body of the helix shows a high preference for aliphatic residues Ala, Leu, Val, and Ile, while these are avoided at helix termini. A propensity scale for amino acids to occur in the middle of helices has been obtained. Comparison of this scale with several previously reported scales shows that this scale correlates best with the experimentally determined values. Proteins 31:460–476, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

16.
The primary objective of this study was to determine whether total biodiversity (γ) is partitioned into within‐community (α) and among‐community (β) components differently for taxonomic and functional organization. I hypothesized that α diversity will contribute more to the functional organization of γ diversity and that β diversity will contribute more to the taxonomic organization of γ diversity. A secondary objective was to determine whether the relationship between taxonomic and functional diversity is scale dependent. Species abundance data was obtained from fisheries surveys conducted by the Texas Parks and Wildlife Dept that focused on least disturbed streams from 11 different ecoregions of Texas, including 62 localities from 18 drainages. Functional and taxonomic organization of assemblages was quantified with two different measures of biodiversity, including richness and the numbers equivalent of Shannon diversity. Scale‐dependent effects on these indices were assessed by multiplicatively partitioning γ into α and β components. The contribution of α and β components to γ diversity differed between functional and taxonomic organization and among different measures of biodiversity. Among‐community components were more influential in structuring the taxonomic organization of stream‐fish assemblages, whereas within‐community components were more important in structuring the functional organization of assemblages. The relationship between taxonomic and functional diversity differed between α and β components and between spatial scales. Indeed, ecological patterns not only change with spatial scale, but how they change is dependent on which aspect of biodiversity is considered.  相似文献   

17.
Phosphorylase kinase (PhK), a 1.3 MDa regulatory enzyme complex in the glycogenolysis cascade, has four copies each of four subunits, (αβγδ)4, and 325 kDa of unique sequence (the mass of an αβγδ protomer). The α, β and δ subunits are regulatory, and contain allosteric activation sites that stimulate the activity of the catalytic γ subunit in response to diverse signaling molecules. Due to its size and complexity, no high resolution structures have been solved for the intact complex or its regulatory α and β subunits. Of PhK's four subunits, the least is known about the structure and function of its largest subunit, α. Here, we have modeled the full‐length α subunit, compared that structure against previously predicted domains within this subunit, and performed hydrogen‐deuterium exchange on the intact subunit within the PhK complex. Our modeling results show α to comprise two major domains: an N‐terminal glycoside hydrolase domain and a large C‐terminal importin α/β‐like domain. This structure is similar to our previously published model for the homologous β subunit, although clear structural differences are present. The overall highly helical structure with several intervening hinge regions is consistent with our hydrogen‐deuterium exchange results obtained for this subunit as part of the (αβγδ)4 PhK complex. Several low exchanging regions predicted to lack ordered secondary structure are consistent with inter‐subunit contact sites for α in the quaternary structure of PhK; of particular interest is a low‐exchanging region in the C‐terminus of α that is known to bind the regulatory domain of the catalytic γ subunit.  相似文献   

18.
Study of the most conserved region in many β/α-barrels, the phosphate-binding site, revealed a sequence motif in a few β/α-barrels with known tertiary structure, namely glycolate oxidase (GOX), cytochrome b2 (Cyb2), tryptophan synthase α subunit (TrpA), and the indoleglycerolphosphate synthase (TrpC). Database searches identified this motif in numerous other enzyme families: (1) IMP dehydrogenase (IMPDH) and GMP reductase (GuaC); (2) phosphoribosylformimino-5-aminoimidazol carboxamide ribotide isomerase (HisA) and the cyclase-producing D-erythro-imidazole-glycerolphosphate (HisF) of the histidine biosynthetic pathway; (3) dihydroorotate dehydrogenase (PyrD); (4) glutamate synthase (GltB); (5) ThiE and ThiG involved in the biosynthesis of thiamine as well as related proteins; (6) an uncharacterized open reading frame from Erwinia herbicola; and (7) a glycerol uptake operon antiterminator regulatory protein (GlpP). Secondary structure predictions of the different families mentioned above revealed an alternating order of β-strands and α-helices in agreement with a β/α-barrel-like topology. The putative phosphate-binding site is always found near the C-terminus of the enzymes, which are all at least about 200 amino acids long. This is compatible with its assumed location between strand 7 and helix 8. The identification of a significant motif in functionally diverse enzymes suggests a divergent evolution of at least a considerable fraction of β/α-barrels. In addition to the known accumulation of β/α-barrels in the tryptophan biosynthetic pathway, we observe clusters of these enzymes in histidine biosynthesis, purine metabolism, and apparently also in thiamine biosynthesis. The substrates are mostly heterocyclic compounds. Although the marginal sequence similarities do not allow a reconstruction of the barrel spreading, they support the idea of pathway evolution by gene duplication.  相似文献   

19.
Aim We investigated partitioning of aquatic macroinvertebrate diversity in eight headwater streams to determine the relative contributions of α and β diversity to γ diversity, and the scale dependence of α and β components. Location Great Dividing Range, Victoria, Australia. Methods We used the method of Jost (Ecology, 2007, 88, 2427–2439) to partition γ diversity into its α and β components. We undertook the analyses at both reach and catchment scales to explore whether inferences depended on scale of observation. Results We hypothesized that β diversity would make a large contribution to the γ diversity of macroinvertebrates in our dendritic riverine landscape, particularly at the larger spatial scale (among catchments) because of limited dispersal among sites and especially among catchments. However, reaches each had relatively high taxon richness and high α diversity, while β diversity made only a small contribution to γ diversity at both the reach and catchment scales. Main conclusions Dendritic riverine landscapes have been thought to generate high β diversity as a consequence of limited dispersal and high heterogeneity among individual streams, but this may not hold for all headwater stream systems. Here, α diversity was high and β diversity low, with individual headwater stream reaches each containing a large portion of γ diversity. Thus, each stream could be considered to have low irreplaceability since losing the option to use one of these sites in a representative reserve network does not greatly diminish the options available for completing the reserve network. Where limited information on individual taxonomic distributions is available, or time and money for modelling approaches are limited, diversity partitioning may provide a useful ‘first‐cut’ for obtaining information about the irreplaceability of individual streams or subcatchments when establishing representative freshwater reserves.  相似文献   

20.
The propensities of peptides that contain the Asn‐Gly segment to form β‐turn and β‐hairpin structures were explored using the density functional methods and the implicit solvation model in CH2Cl2 and water. The populations of preferred β‐turn structures varied depending on the sequence and solvent polarity. In solution, β‐hairpin structures with βI′ turn motifs were most preferred for the heptapeptides containing the Asn‐Gly segment regardless of the sequence of the strands. These preferences in solution are consistent with the corresponding X‐ray structures. The sequence, H‐bond strengths, solvent polarity, and conformational flexibility appeared to interact to determine the preferred β‐hairpin structure of each heptapeptide, although the β‐turn segments played a role in promoting the formation of β‐hairpin structures and the β‐hairpin propensity varied. In the heptapeptides containing the Asn‐Gly segment, the β‐hairpin formation was enthalpically favored and entropically disfavored at 25°C in water. The calculated results for β‐turns and β‐hairpins containing the Asn‐Gly segment imply that these structural preferences may be useful for the design of bioactive macrocyclic peptides containing β‐hairpin mimics and the design of binding epitopes for protein–protein and protein–nucleic acid recognitions. © 2016 Wiley Periodicals, Inc. Biopolymers 105: 653–664, 2016.  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