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.
We present a fully automatic structural classification of supersecondary structure units, consisting of two hydrogen-bonded β strands, preceded or followed by an α helix. The classification is performed on the spatial arrangement of the secondary structure elements, irrespective of the length and conformation of the intervening loops. The similarity of the arrangements is estimated by a structure alignment procedure that uses as similarity measure the root mean square deviation of superimposed backbone atoms. Applied to a set of 141 well-resolved nonhomologous protein structures, the classification yields 11 families of recurrent arrangements. In addition, fragments that are structurally intermediate between the families are found; they reveal the continuity of the classification. The analysis of the families shows that the α helix and β hairpin axes can adopt virtually all relative orientations, with, however, some preferable orientations; moreover, according to the orientation, preferences in the left/right handedness of the α–β connection are observed. These preferences can be explained by favorable side by side packing of the α helix and the β hairpin, local interactions in the region of the α–β connection or stabilizing environments in the parent protein. Furthermore, fold recognition procedures and structure prediction algorithms coupled to database-derived potentials suggest that the preferable nature of these arrangements does not imply their intrinsic stability. They usually accommodate a large number of sequences, of which only a subset is predicted to stabilize the motif. The motifs predicted as stable could correspond to nuclei formed at the very beginning of the folding process. Proteins 30:193–212, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

2.
In anterior pituitaries from male rats, it appeared that 5α-androstane-3β, 17β-diol was quickly metabolized into 5α-androstane-3β,6α-17β-triol and 5α-androstane-3β,7α, 17β-triol by action of 6α- and 7α-hydroxylases. Hydroxysteroid hydroxylases were located in endoplasmic reticulum and were dependent on NADPH+. Their optimum pH was 8.0, optima temperature, 37°C, and their apparent Km was 2.7 μM. Hydroxylative reactions were not reversible and not modified by gonadectomy. Hydroxylation seemed an efficient control of the pituitary level of 5α-andros-tane-3β, 17β-diol.  相似文献   

3.
4.
5.
The past two decades have seen great progress in understanding the mechanisms of ecosystem stability in local ecological systems. There is, however, an urgent need to extend existing knowledge to larger spatial scales to match the scale of management and conservation. Here, we develop a general theoretical framework to study the stability and variability of ecosystems at multiple scales. Analogously to the partitioning of biodiversity, we propose the concepts of alpha, beta and gamma variability. Gamma variability at regional (metacommunity) scale can be partitioned into local alpha variability and spatial beta variability, either multiplicatively or additively. On average, variability decreases from local to regional scales, which creates a negative variability–area relationship. Our partitioning framework suggests that mechanisms of regional ecosystem stability can be understood by investigating the influence of ecological factors on alpha and beta variability. Diversity can provide insurance effects at the various levels of variability, thus generating alpha, beta and gamma diversity–stability relationships. As a consequence, the loss of biodiversity and habitat impairs ecosystem stability at the regional scale. Overall, our framework enables a synthetic understanding of ecosystem stability at multiple scales and has practical implications for landscape management.  相似文献   

6.
The crystal state conformations of three peptides containing the α,α-dialkylated residues. α,α-di-n-propylglycine (Dpg) and α,α-di-n-butylglycine (Dbg), have been established by x-ray diffraction. Boc-Ala-Dpg-Alu-OMe (I) and Boc-Ala-Dbg-Ala-OMe (III) adopt distorted type II β-turn conformations with Ala (1) and Dpg/Dbg (2) as the corner residues. In both peptides the conformational angles at the Dxg residue (I: ? = 66.2°, ψ = 19.3°; III: ? = 66.5°. ψ = 21.1°) deviate appreciably from ideal values for the i + 2 residue in a type II β-turn. In both peptides the observed (N…O) distances between the Boc CO and Ala (3) NH groups are far too long (1: 3.44 Å: III: 3.63 Å) for an intramolecular 4 → 1 hydrogen bond. Boc-Ala-Dpg-Ata-NHMe (II) crystallizes with two independent molecules in the asymmetric unit. Both molecules HA and HB adopt consecutive β-turn (type III-III in HA and type III-I in IIB) or incipient 310-helical structures, stabilized by two intramolecular 4 → 1 hydrogen bonds. In all four molecules the bond angle N-Cα-C′ (τ) at the Dxg residues are ≥ 110°. The observation of conformational angles in the helical region of ?,ψ space at these residues is consistent with theoretical predictions. © 1995 John Wiley & Sons, Inc.  相似文献   

7.
Natural vitamin E consists of four different tocopherol and four different tocotrienol homologues (α, β, γ, δ) that all have antioxidant activity. However, recent data indicate that the different vitamin E homologues also have biological activity unrelated to their antioxidant activity. In this review, we discuss the anti-inflammatory properties of the two major forms of vitamin E, α-tocopherol (αT) and γ-tocopherol (γT), and discuss the potential molecular mechanisms involved in these effects. While both tocopherols exhibit anti-inflammatory activity in vitro and in vivo, supplementation with mixed (γT-enriched) tocopherols seems to be more potent than supplementation with αT alone. This may explain the mostly negative outcomes of the recent large-scale interventional chronic disease prevention trials with αT only and thus warrants further investigation.  相似文献   

8.
Tuberculosis is still affecting millions of people worldwide, and new resistant strains of Mycobacterium tuberculosis are being found. It is therefore necessary to find new compounds for treatment. In this paper, we report the synthesis and in vitro testing of peptidyl β‐aminoboronic acids and β‐aminoboronates with anti‐tubercular activity. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
There is a critical need for compounds that target cell surface integrin receptors for applications in cancer therapy and diagnosis. We used directed evolution to engineer the Ecballium elaterium trypsin inhibitor (EETI‐II), a knottin peptide from the squash family of protease inhibitors, as a new class of integrin‐binding agents. We generated yeast‐displayed libraries of EETI‐II by substituting its 6‐amino acid trypsin binding loop with 11‐amino acid loops containing the Arg‐Gly‐Asp integrin binding motif and randomized flanking residues. These libraries were screened in a high‐throughput manner by fluorescence‐activated cell sorting to identify mutants that bound to αvβ3 integrin. Select peptides were synthesized and were shown to compete for natural ligand binding to integrin receptors expressed on the surface of U87MG glioblastoma cells with half‐maximal inhibitory concentration values of 10–30 nM. Receptor specificity assays demonstrated that engineered knottin peptides bind to both αvβ3 and αvβ5 integrins with high affinity. Interestingly, we also discovered a peptide that binds with high affinity to αvβ3, αvβ5, and α5β1 integrins. This finding has important clinical implications because all three of these receptors can be coexpressed on tumors. In addition, we showed that engineered knottin peptides inhibit tumor cell adhesion to the extracellular matrix protein vitronectin, and in some cases fibronectin, depending on their integrin binding specificity. Collectively, these data validate EETI‐II as a scaffold for protein engineering, and highlight the development of unique integrin‐binding peptides with potential for translational applications in cancer. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

10.
α7β1 is the major integrin complex expressed in differentiated muscle cells where it functions as a laminin receptor. In this work we have expressed the α7 integrin subunit in CHO cells to investigate the functional properties of this receptor. After transfection with α7 CHO cells acquired the ability to adhere and spread on laminin 1 consistent with the laminin receptor activity of the α7β1. α7 transfectants, however, showed a 70% reduction in the ability to adhere to fibronectin and were unable to assemble a fibronectin matrix. The degree of reduction was inversely related to the level of α7 expression. To define the mechanisms underlying this adhesive defect we analyzed surface expression and functional properties of the α5β1 fibronectin receptor. Although cell surface expression of α5β1 was reduced by a factor of 20–25% in α7 transfectants compared to control untransfected cells, this slight reduction was not sufficient to explain the dramatic reduction in cell adhesion (70%) and matrix assembly (close to 100%). Binding studies showed that the affinity of125I-fibronectin for its surface receptor was decreased by 50% in α7 transfectants, indicating that the α5β1 integrin is partially inactivated in these cells. Inactivation can be reversed by Mn2+, a cation known to increase integrin affinity for their ligands. In fact, incubation of cells with Mn2+restored fibronectin binding affinity, adhesion to fibronectin, and assembly of fibronectin matrix in α7 transfectants. These data indicate that α7 expression leads to the functional down regulation of α5β1 integrin by decreasing ligand binding affinity and surface expression. In conclusion, the data reported establish the existence of anegative cooperativitybetween α7 and α5 integrins that may be important in determining functional regulation of integrins during myogenic differentiation.  相似文献   

11.
5α‐Androst‐16‐en‐3α‐ol (α‐androstenol) is an important contributor to human axilla sweat odor. It is assumed that α‐andostenol is excreted from the apocrine glands via a H2O‐soluble conjugate, and this precursor was formally characterized in this study for the first time in human sweat. The possible H2O‐soluble precursors, sulfate and glucuronide derivatives, were synthesized as analytical standards, i.e., α‐androstenol, β‐androstenol sulfates, 5α‐androsta‐5,16‐dien‐3β‐ol (β‐androstadienol) sulfate, α‐androstenol β‐glucuronide, α‐androstenol α‐glucuronide, β‐androstadienol β‐glucuronide, and α‐androstenol β‐glucuronide furanose. The occurrence of α‐androstenol β‐glucuronide was established by ultra performance liquid chromatography (UPLC)/MS (heated electrospray ionization (HESI)) in negative‐ion mode in pooled human sweat, containing eccrine and apocrine secretions and collected from 25 female and 24 male underarms. Its concentration was of 79 ng/ml in female secretions and 241 ng/ml in male secretions. The release of α‐androstenol was observed after incubation of the sterile human sweat or α‐androstenol β‐glucuronide with a commercial glucuronidase enzyme, the urine‐isolated bacteria Streptococcus agalactiae, and the skin bacteria Staphylococcus warneri DSM 20316, Staphylococcus haemolyticus DSM 20263, and Propionibacterium acnes ATCC 6919, reported to have β‐glucuronidase activities. We demonstrated that if α‐ and β‐androstenols and androstadienol sulfates were present in human sweat, their concentrations would be too low to be considered as potential precursors of malodors; therefore, the H2O‐soluble precursor of α‐androstenol in apocrine secretion should be a β‐glucuronide.  相似文献   

12.
As the prevalence of osteoporosis is expected to increase over the next few decades, the development of novel therapeutic strategies to combat this disorder becomes clinically imperative. These efforts draw extensively from an expanding body of knowledge pertaining to the physiologic mechanisms of skeletal homeostasis. To this body of knowledge, we contribute that cells of hematopoietic lineage may play a crucial role in balancing osteoblastic bone formation against osteoclastic resorption. Specifically, our laboratory has previously demonstrated that megakaryocytes (MKs) can induce osteoblast (OB) proliferation in vitro, but do so only when direct cell‐to‐cell contact is permitted. To further investigate the nature of this interaction, we have effectively neutralized several adhesion molecules known to function in the analogous interaction of MKs with another cell type of mesenchymal origin—the fibroblast (FB). Our findings implicate the involvement of fibronectin/RGD‐binding integrins including α3β1 (VLA‐3) and α5β1 (VLA‐5) as well as glycoprotein (gp) IIb (CD41), all of which are known to be expressed on MK membranes. Furthermore, we demonstrate that interleukin (IL)‐3 can enhance MK‐induced OB activation in vitro, as demonstrated in the MK–FB model system. Taken together, these results suggest that although their physiologic and clinical implications are very different, these two models of hematopoietic–mesenchymal cell activation are mechanistically analogous in several ways. J. Cell. Biochem. 109: 927–932, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

13.
One of the hallmarks of cells undergoing mitotic division is their rounded morphology and reduced adhesion to the substratum. We have studied and compared the attachment of interphase and mitotic cells to substrata coated with fibronectin and vitronectin. We have found that adhesion of mitotic cells, as compared to interphase cells, is significantly reduced to fibronectin, but is higher to vitronectin. These results correlate well with the expression of α5β1 and αVβ3 integrins, the respective receptors for fibronectin and vitronectin, on the cell surface. Mitotic cells show higher levels of αVβ3 and very low levels of α5β1 proteins on the cell surface as compared to interphase cells. This difference in the levels of these integrins also reflects in the total amounts of fibronectin and vitronectin present on the cell surface of these cells. We have further shown, by flow cytometry, that binding of vitronectin, or the synthetic peptide-GRGDSP-, causes an increase in the intracellular levels of Ca2− in mitotic cells, but no change is seen in the interphase cells. Binding of fibronectin to either of these cells fails to elicit any response. One interesting feature of our results is that the levels of total, i.e., cytoplasmic plus membrane bound, α5β1 and αVβ3 integrins of mitotic and interphase cells remain the same, thus implying an alteration in the distribution of integrin chains between the plasma membrane and the cytoplasm during the conversion of interphase cells into the mitotic phase. © 1996 Wiley-Liss, Inc.  相似文献   

14.
Laminins, a family of heterotrimeric proteins with cell adhesive/signaling properties, are characteristic components of basement membranes of vasculature and tissues. In the present study, permeabilized platelets were found to react with a monoclonal antibody to laminin γ1 chain by immunofluorescence. In Western blot analysis of platelet lysates, several monoclonal antibodies to γ1 and β1 laminin chains recognized 220- to 230-kDa polypeptides, under reducing conditions, and a structure with much slower electrophoretic mobility under nonreducing conditions. Immunoaffinity purification on a laminin β1 antibody–Sepharose column yielded polypeptides of 230, 220, 200, and 180 kDa from platelet lysates. In the purified material, mAbs to β1 and γ1 reacted with the two larger polypeptides, while affinity-purified rabbit antibodies to laminin α4 chain recognized the smallest polypeptide. Identity of the polypeptides was confirmed by microsequencing. One million platelets contained on average 1 ng of laminin (approximately 700 molecules per cell), of which 20–35% was secreted within minutes after stimulation with either thrombin or phorbol ester. Platelets adhered to plastic surfaces coated with the purified platelet laminin, and this process was largely inhibited by antibodies to β1 and α6 integrin chains. We conclude that platelets contain and, following activation, secrete laminin-8 (α4β1γ1) and that the cells adhere to the protein by using α6β1 integrin.  相似文献   

15.
In the present paper we describe the synthesis, purification, single-crystal x-ray analysis, solution conformational characterization, and conformational energy calculations of the cyclic tetrapeptide cyclo- (β-Ala-L -Pro-β-Ala-L -Val). The peptide was synthesized by classical solution methods and the cyclization of the free tetrapeptide was accomplished in good yields in diluted methylene chloride solution using N,N-dicyclohexyl-carbodiimide. The compound crystallizes in the monoclinic space group P21 from ethanol with two independent molecules in the unit cell. All peptide bonds are trans. The nmr molecular conformation in the acetonitrile solution as well as that derived from the molecular dynamic simulation in vacuo is quite different from those observed in the solid state and is very similar to that previously observed for the parent compound cyclo-(β-Ala-L -Pro-β-Ala-L -Pro). © 1993 John Wiley & Sons, Inc.  相似文献   

16.
β‐dystroglycan (β‐DG) is a widely expressed transmembrane protein that plays important roles in connecting the extracellular matrix to the cytoskeleton, and thereby contributing to plasma membrane integrity and signal transduction. We previously observed nuclear localization of β‐DG in cultured cell lines, implying the existence of a nuclear targeting mechanism that directs it to the nucleus instead of the plasma membrane. In this study, we delineate the nuclear import pathway of β‐DG, characterizing a functional nuclear localization signal (NLS) in the β‐DG cytoplasmic domain, within amino acids 776–782. The NLS either alone or in the context of the whole β‐DG protein was able to target the heterologous GFP protein to the nucleus, with site‐directed mutagenesis indicating that amino acids R779 and K780 are critical for NLS functionality. The nuclear transport molecules Importin (Imp)α and Impβ bound with high affinity to the NLS of β‐DG and were found to be essential for NLS‐dependent nuclear import in an in vitro reconstituted nuclear transport assay; cotransfection experiments confirmed the dependence on Ran for nuclear accumulation. Intriguingly, experiments suggested that tyrosine phosphorylation of β‐DG may result in cytoplasmic retention, with Y892 playing a key role. β‐DG thus follows a conventional Impα/β‐dependent nuclear import pathway, with important implications for its potential function in the nucleus. J. Cell. Biochem. 110: 706–717, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

17.
Microtubules provide structural support for a cell and play key roles in cell motility, mitosis, and meiosis. They are also the targets of several anticancer agents, indicating their importance in maintaining cell viability. We have investigated the possibility that alterations in microtubule structure and tubulin polymerization may be part of the cellular response to DNA damage. In this report, we find that γ-radiation stimulates the production and polymerization of α-, β-, and γ- tubulin in hematopoeitic cell lines (Ramos, DP16), leading to visible changes in microtubule structures. We have found that this microtubule reorganization can be prevented by caffeine, a drug that concomitantly inhibits DNA damage-induced cell cycle arrest and apoptosis. Our results support the idea that microtubule polymerization is an important facet of the mammalian response to DNA damage.  相似文献   

18.
During tissue morphogenesis and tumor invasion, epithelial cells must undergo intercellular rearrangement in which cells are repositioned with respect to one another and the surrounding mesenchymal extracellular matrix. Using three-dimensional aggregates of squamous epithelial cells, we show that such intercellular rearrangements can be triggered by activation of β1 integrins after their ligation with extracellular matrices. On nonadherent substrates, multicellular aggregates (MCAs) formed rapidly via E-cadherin junctional complexes and over time became compacted spheroids exhibiting a more epithelial phenotype. After MCAs were replated on culture substrates, the spheroids collapsed to yield tightly arranged cell monolayers. Cell–cell contact induced rapid elevation in E-cadherin levels, which was due to an increase in the metabolic stability of junctional receptors. During MCA remodeling of cell–cell adhesions, and monolayer formation, their E-cadherin levels fell rapidly. Similar behavior was obtained regardless of which ECM ligand—collagen type I, fibronectin, or laminin 1—MCAs were seeded on. In contrast, when seeded onto a matrix elaborated by squamous epithelial cells, cells in the MCA attached, spread, lost cell–cell junctions, and dispersed. Analysis identified laminin 5 as the active ECM ligand in this matrix, and MCA dispersion required functional β1 integrin and specifically α3β1. Furthermore, substrate-immobilized anti-integrin antibody effectively reproduced the epithelial–mesenchymal-like transition induced by the laminin 5 matrix. During the early stages of aggregate rearrangement and collapse, cells on laminin 5 substrates, but not those on collagen I substrates, exhibited intense cortical arrays of F-actin, microspikes, and fascin accumulation at their peripheral surfaces. These results suggest that engagement of specific integrin–ligand pairs regulates cadherin junctional adhesions during events common to epithelial morphogenesis and tumor invasion.  相似文献   

19.
The conformational preferences of linear peptides containing α,α-disubstituted α-amino acids, derived from the crystal structures of 28 compounds, are reviewed. In particular, the sensitivity of peptide conformation to the geometry of these unusual amino acids is underlined. We also consider possible future directions of research, which, we hope, will result in a complete understanding of the structures adopted by peptaibol antibiotics.  相似文献   

20.
Type 3 17β-hydroxysteroid dehydrogenase (17β-HSD), a key steroidogenic enzyme, transforms 4-androstene-3,17-dione (Δ4-dione) into testosterone. In order to produce potential inhibitors, we performed solid-phase synthesis of model libraries of 3β-peptido-3α-hydroxy-5α-androstan-17-ones with 1, 2, or 3 levels of molecular diversity, obtaining good overall yields (23–58%) and a high average purity (86%, without any purification steps) using the Leznoff's acetal linker. The libraries were rapidly synthesized in a parallel format and the generated compounds were tested as inhibitors of type 3 17β-HSD. Potent inhibitors were identified from these model libraries, especially six members of the level 3 library having at least one phenyl group. One of them, the 3β-(N-heptanoyl- -phenylalanine- -leucine-aminomethyl)-3α-hydroxy-5α-androstan-17-one (42) inhibited the enzyme with an IC50 value of 227 nM, which is twice as potent as the natural substrate Δ4-dione when used itself as an inhibitor. Using the proliferation of androgen-sensitive (AR+) Shionogi cells as model of androgenicity, the compound 42 induced only a slight proliferation at 1 μM (less than previously reported type 3 17β-HSD inhibitors) and, interestingly, no proliferation at 0.1 μM.  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