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.
During platelet activation, fibrinogen binds to its specific platelet receptor, integrin αIIbβ3, thus completing the final common pathway for platelet aggregation. Norcantharidin (NCTD) is a promising anticancer agent in China from medicinal insect blister beetle. In this study, we provided the evidence to demonstrate NCTD (0.1–1.0 μM) possesses very powerful antiplatelet activity in human platelets; nevertheless, it had no effects on surface P‐selectin expression and only slight inhibition on ATP‐release reaction in activated platelets. Moreover, NCTD markedly hindered integrin αIIbβ3 activation by interfering with the binding of FITC‐labelled PAC‐1. It also markedly reduced the number of adherent platelets and the single platelet spreading area on immobilized fibrinogen as well as clot retraction. Additionally, NCTD attenuated phosphorylation of proteins such as integrin β3, Src and FAK in platelets spreading on immobilized fibrinogen. These results indicate that NCTD restricts integrin αIIbβ3‐mediated outside‐in signalling in human platelets. Besides, NCTD substantially prolonged the closure time in human whole blood and increased the occlusion time of thrombotic platelet plug formation and prolonged the bleeding time in mice. In conclusion, NCTD has dual activities, it can be a chemotherapeutic agent for cancer treatment, and the other side it possesses powerful antiplatelet activity for treating thromboembolic disorders.  相似文献   

2.
We report here an analysis of the expression and function of the α chain of human VLA-4 in stable mouse L cell transfectants and the requirement for the β chain in these processes. L cells were transfected with human α4 cDNA or α4 and human β1 cDNA. Unexpectedly, human α4 cDNA, when transfected alone, could induce de novo surface expression of host β7 and increased expression of host β1. Induction of mouse β7 and β1 surface expression was not due to de novo gene activation, but instead represented α4/β intracellular subunit association and transport to the cell surface. Transfection with human β1 prevented surface expression of mouse β integrins. Whereas human α4 and human β1 subunits associated very tightly in anti-α4 immunoprecipitates, human α4 and mouse β subunits were only partially associated. Furthermore, binding of human/mouse chimeric receptors to recombinant VCAM, a major ligand for α4β7 and α4β1, was very poor, whereas human α4/human β1 receptors bound strongly to VCAM. One α4 transfectant, which exhibited a tight human α4/mouse β1 association, could be induced, but only after PMA activation, to bind strongly to VCAM. These results indicate that α4 subunits have specific affinity for β7 and β1 integrins and require β subunits for surface expression as well as high affinity ligand binding activity. Our results indicate that a tight association between the α4 and β subunit appears to be critical for ligand binding, consistent with a direct as well as regulatory role for the β subunit in ligand binding. Furthermore, these studies demonstrate that expression of foreign recombinant proteins can alter host cell protein expression resulting in de novo surface protein expression. © 1996 Wiley-Liss, Inc.  相似文献   

3.
The integrin α4β1 is involved in mediating exfiltration of leukocytes from the vasculature. It interacts with a number of proteins up-regulated during the inflammatory response including VCAM-1 and the CS-1 alternatively spliced region of fibronectin. In addition it binds the multifunctional protein osteopontin (OPN), which can act as both a cytokine and an extracellular matrix molecule. Here we map the region of human OPN that supports cell adhesion via α4β1 using GST fusion proteins. We show that α4β1 expressed in J6 cells interacts with intact OPN when the integrin is in a high activation state, and by deletion mapping that the α4β1 binding region in OPN lies between amino acid residues 125 and 168 (aa125–168). This region contains the central RGD motif of OPN, which also interacts with integrins αvβ3, αvβ5, αvβ1, α8β1, and α5β1. Mutating the RGD motif to RAD had no effect on the interaction with α4β1. To define the binding site the region incorporating aa125–168 was divided into 5 overlapping peptides expressed as GST fusion proteins. Two peptides supported adhesion via α4β1, aa132–146, and aa153–168; of these only a synthetic peptide, SVVYGLR (aa162–168), derived from aa153–168 was able to inhibit α4β1 binding to CS-1. These data identify the motif SVVYGLR as a novel peptide inhibitor of α4β1, and the primary α4β1 binding site within OPN.  相似文献   

4.
3α-Hydroxysteroid dehydrogenases (3α-HSDs) inactivate steroid hormones in the liver, regulate 5α-dihydrotestosterone (5α-DHT) levels in the prostate, and form the neurosteroid, allopregnanolone in the CNS. Four human 3α-HSD isoforms exist and correspond to AKR1C1–AKR1C4 of the aldo-keto reductase (AKR) superfamily. Unlike the related rat 3α-HSD (AKR1C9) which is positional and stereospecific, the human enzymes display varying ratios of 3-, 17-, and 20-ketosteroid reductase activity as well as 3α-, 17β-, and 20α-hydroxysteroid oxidase activity. Their kcat values are 50–100-fold lower than that observed for AKR1C9. Based on their product profiles and discrete tissue localization, the human enzymes may regulate the levels of active androgens, estrogens, and progestins in target tissues. The X-ray crystal structures of AKR1C9 and AKR1C2 (human type 3 3α-HSD, bile acid binding protein and peripheral 3α-HSD) reveal that the AKR1C2 structure can bind steroids backwards (D-ring in the A-ring position) and upside down (β-face inverted) relative to the position of a 3-ketosteroid in AKR1C9 and this may account for its functional plasticity. Stopped-flow studies on both enzymes indicate that the conformational changes associated with binding cofactor (the first ligand) are slow; they are similar in both enzymes but are not rate-determining. Instead the low kcat seen in AKR1C2 (50-fold less than AKR1C9) may be due to substrate “wobble” at the plastic active site.  相似文献   

5.
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   

6.
The oncogenic SHC proteins are signaling substrates for most receptor and cytoplasmic tyrosine kinases (TKs) and have been implicated in cellular growth, transformation, and differentiation. In tumor cells overexpressing TKs, the levels of tyrosine phosphorylated SHC are chronically elevated. The significance of amplified SHC signaling in breast tumorigenesis and metastasis remains unknown. Here we demonstrate that seven- to ninefold overexpression of SHC significantly altered interactions of cells with fibronectin (FN). Specifically, in human breast cancer cells overexpressing SHC (MCF-7/SHC) the association of SHC with α5β1 integrin (FN receptor) was increased, spreading on FN was accelerated, and basal growth on FN was reduced. These effects coincided with an early decline of adhesion-dependent MAP kinase activity. Basal motility of MCF-7/SHC cells on FN was inhibited relative to that in several cell lines with normal SHC levels. However, when EGF or IGF-I was used as the chemoattractant, the locomotion of MCF-7/SHC cells was greatly (approx fivefold) stimulated, while it was only minimally altered in the control cells. These data suggest that SHC is a mediator of the dynamic regulation of cell adhesion and motility on FN in breast cancer cells.  相似文献   

7.
How to target cancer cells with high specificity and kill cancer cells with high efficiency remains an urgent demand for anticancer drugs. Temporin‐La, which belongs to the family of temporins, presents antitumor activity against many cancer cell lines. We first used a whole bioinformatic analysis method as a platform to identify new anticancer antimicrobial peptides (AMPs). On the basis of these results, we designed a temporin‐La analog (temporin‐Las) and related constructs containing the Arg‐Gly‐Asp (RGD) tripeptide, the integrin αvβ3 homing domain (RGD‐La and RGD‐Las). We detected a link between the net charges and integrin αvβ3 expression of cancer cell lines and the antitumor activities of these peptides. Temporin‐La and its synthetic analogs inhibited cancer cell proliferation in a dose‐dependent manner. Evidence was provided that the affinity between RGD‐Las and tumor cell membranes was stronger than other tested peptides using a pull‐down assay. Morphological changes on the cell membrane induced by temporin‐La and RDG‐Las, respectively, were examined by scanning electron microscopy. Additionally, time‐dependent morphological changes were detected by confocal microscopy, where the binding process of RGD‐Las to the cell membrane could be monitored. The results indicate that the electrostatic interaction between these cationic peptides and the anionic cell membrane is a major determinant of selective cell killing. Thus, the RGD tripeptide is a valuable ligand motif for tumor targeting, which leads to an increased anticancer efficiency by RGD‐Las. These AMP‐derived peptides have clinical potential as specifically targeting agents for the treatment of αvβ3 positive tumors. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
β‐Amino acids containing hybrid peptides and β‐peptides show great potential as peptidomimetics. In this paper we describe the synthesis and affinity toward the µ‐ and δ‐opioid receptors of β‐peptides, analogues of Leu‐enkephalin, deltorphin I, dermorphin and α,β‐hybrides, analogues of deltorphin I. Substitution of α‐amino acid residues with β3homo‐amino acid residues, in general resulted in decrease of affinity to opioid receptors. However, the incorporation β3h‐D ‐Ala in position 2 or β3hPhe in position 3 of deltorphin I resulted in potent and selective ligand for δ‐opioid receptor. The NMR studies of β‐deltorphin I analogue suggest that conformational motions in the central part of the peptide backbone are partially restricted and some conformational preferences can be expected. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
The present study describes the synthesis and crystallographic analysis of αγ hybrid peptides, Boc‐Gpn‐L‐Pro‐NHMe ( 1 ), Boc‐Aib‐Gpn‐L‐Pro‐NHMe ( 2 ), and Boc‐L‐Pro‐Aib‐Gpn‐L‐Pro‐NHMe ( 3 ). Peptides 1 and 2 adopt expanded 12‐membered (C12) helical turn over γα segment. Peptide 3 promotes the ribbon structure stabilized by type II β‐turn (C10) followed by the expanded C12 helical γα turn. Both right‐handed and left‐handed helical conformations for Aib residue are observed in peptides 2 and 3 , respectively Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
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   

11.
12.
Hematopoietic homeostasis depends on the maintenance of hematopoietic stem cells (HSCs), which are regulated within a specialized bone marrow (BM) niche. When HSC sense external stimuli, their adhesion status may be critical for determining HSC cell fate. The cell surface molecule, integrin αvβ3, is activated through HSC adhesion to extracellular matrix and niche cells. Integrin β3 signaling maintains HSCs within the niche. Here, we showed the synergistic negative regulation of the pro‐inflammatory cytokine interferon‐γ (IFNγ) and β3 integrin signaling in murine HSC function by a novel definitive phenotyping of HSCs. Integrin αvβ3 suppressed HSC function in the presence of IFNγ and impaired integrin β3 signaling mitigated IFNγ‐dependent negative action on HSCs. During IFNγ stimulation, integrin β3 signaling enhanced STAT1‐mediated gene expression via serine phosphorylation. These findings show that integrin β3 signaling intensifies the suppressive effect of IFNγ on HSCs, which indicates that cell adhesion via integrin αvβ3 within the BM niche acts as a context‐dependent signal modulator to regulate the HSC function under both steady‐state and inflammatory conditions.  相似文献   

13.
14.
d ‐β‐aspartyl (Asp) residue has been found in a living body such as aged lens crystallin, although l ‐α‐amino acids are constituents in natural proteins. Isomerization from l ‐α‐ to d ‐β‐Asp probably modulates structures to affect biochemical reactions. At Asp residue, isomerization and peptide bond cleavage compete with each other. To gain insight into how fast each reaction proceeds, the analysis requires the consideration of both pathways simultaneously and independently. No information has been provided, however, about these competitive processes because each reaction has been studied separately. The contribution of Asp isomers to the respective pathways has still been veiled. In this work, the two competitive reactions, isomerization and spontaneous peptide bond cleavage at Asp residue, were simultaneously observed and compared in an αA‐crystallin fragment, S51LFRTVLD58SG60 containing l ‐α‐ and d ‐β‐Asp58 isomers. The kinetics showed that the formation of l ‐ and d ‐succinimide (Suc) intermediate, as a first step of isomerization, was comparable at l ‐α‐ and d ‐β‐Asp. Although l ‐Suc was converted to l ‐β‐Asp, d ‐Suc was liable to return to the original d ‐β‐Asp, the reverse reaction marked enough to consider d ‐β‐Asp as apparently stable. d ‐β‐Asp was also resistant to the peptide bond cleavage. Such apparent less reactivity is probably the reason for gradual and abnormal accumulation of d ‐β‐Asp in a living body under physiological conditions. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
16.
Simulated enzymic debranching of a β-limit dextrin model, prepared from a computed construct made by random extension and branching, and given the CCL value of w-maize amylopectin (and equal amounts of external chains with ECL values of 2 and 3) has been related to experimental chromatograms of the debranched β-limit dextrin of the amylopectin. The profile was similar to those from gel chromatograms and IEC-PAD chromatography.The equivalent lengths in glucosyl units of grid-links (g-links) of internal and external chains in constructs were calculated from the ICL and ECL values of amylopectin and models produced from the constructs with the appropriate lengths for internal and external chains. These derived models were subjected to simulated hydrolysis by Pseudomonas stutzeri amylase and the products compared with those of the experimental distribution from w-maize amylopectin. With the model the amounts of maltotetraose and maltodextrins released were similar to the experimental values but the distribution of branched maltodextrins was quite different. Unlike w-maize amylopectin – a polymer with the cluster structure – which has given a profile of molecular sizes of maltodextrins with low amounts of single and small numbers of internal chains and with a peak at a MW of about 14,000 (13 chains), in the model the proportion of maltodextrin with one internal chain was high and as d.p. increased the amounts decreased exponentially. This would be expected if the distribution of internal chains in the core was random. It is suggested that in the core of a model prepared from a construct made with alternating probabilities of extension – one in which this probability is high relative to branching, and a second in which it is low – may give clusters of branched maltodextrins with short internal chains which are joined by longer chains; more closely approximating the distribution of internal chains of different lengths in amylopectin.An arrangement for amylopectin molecules in the starch granule has been proposed. In this, they have a wafer-like, discoidal shape, composed of the amorphous zone overlain with the double helical, crystalline region. The flat macromolecules are concentrically layered with the former on the inside and the latter oriented to the outside of the granule.  相似文献   

17.
Previous reports indicated that integrins associated signals are tightly related to tumor progression. Here, we observed elevated expression of integrin α2β1 in tumor tissues from microtubule‐directed chemotherapeutic drugs (MDCDs) resistant patients compared with the samples from chemosensitive patients. More importantly, we sorted the integrin α2β1+ tumor cells and found those cells revealed high MDCDs resistance, whereas MDCDs shows effective cytotoxicity to those integrin α2β1? tumor cells in vitro and in vivo. Mechanistically, we demonstrated that integrin α2β1 could induce MDCDs resistance through the activation of the PI3K/AKT pathway. Applying MPEG‐PLA to co‐encapsulate the integrin α2β1 inhibitor E7820 and MDCDs could effectively reverse MDCDs resistance, resulting in enhanced anticancer effects while avoiding potential systemic toxicity in vitro and in vivo. In conclusion, the expression of integrin α2β1 contributes to MDCDs resistance, while applying E7820 combination treatment by MPEG‐PLA nanoparticles could reverse the resistance.  相似文献   

18.
Complement fragment iC3b serves as a major opsonin for facilitating phagocytosis via its interaction with complement receptors CR3 and CR4, also known by their leukocyte integrin family names, αMβ2 and αXβ2, respectively. Although there is general agreement that iC3b binds to the αM and αX I-domains of the respective β2-integrins, much less is known regarding the regions of iC3b contributing to the αX I-domain binding. In this study, using recombinant αX I-domain, as well as recombinant fragments of iC3b as candidate binding partners, we have identified two distinct binding moieties of iC3b for the αX I-domain. They are the C3 convertase-generated N-terminal segment of the C3b α’-chain (α’NT) and the factor I cleavage-generated N-terminal segment in the CUBf region of α-chain. Additionally, we have found that the CUBf segment is a novel binding moiety of iC3b for the αM I-domain. The CUBf segment shows about a 2-fold higher binding activity than the α’NT for αX I-domain. We also have shown the involvement of crucial acidic residues on the iC3b side of the interface and basic residues on the I-domain side.  相似文献   

19.
By introducing an α3 gene-containing plasmid into a human T cell line Jurkat, we prepared the T cells, which express a high level of the α3β1 integrin, to assess the role of laminin 5 in the skin immune system. The α3β1-expressing T cells adhered to laminin 5 and exhibited spreading. These adhered T cells showed a significant tyrosine phosphorylation of intracellular proteins including p59fynupon T-cell receptor (TCR) stimulation. Six hours after cross-linking TCR, these cells on laminin 5 secreted a three times higher level of IL-2 than those on a BSA-coated plate. Twenty hours after the stimulation, 48% of the α3β1-expressing T cells on laminin 5 caused apoptosis. The protein level of cyclin D3 and E decreased, while that of p53 increased in these T cells. These data suggest that laminin 5 may play at least two regulatory roles for T cell functions: augmentation of IL-2 production by antigen-stimulated T cells and induction of apoptosis in these T cells.  相似文献   

20.
Prostaglandin H synthase (PHS) has gained interest as a drugmetabolizing enzyme and has been shown to cooxidize and metabolically activate diethylstilbestrol (DES) in vitro. Both 7,8-benzoflavone (α-naphthoflavone, ANF) and 5,6-benzoflavone (β-naphthoflavone, BNF) have now been studied for their effects on PHS from ram seminal vesicle microsomes by means of several in vitro assays. The PHS-catalyzed cooxidation of DES, as measured by high-performance liquid chromatography (HPLC) analysis, is inhibited by BNF and ANF at micromolar concentrations, with median inhibitory concentrations (IC-50) of<20 and 40 μM, respectively. The oxidation of DES is inhibited whether it is initiated by arachidonic acid or by hydrogen peroxide, indicating that the benzoflavones inhibit PHS by a mechanism different from that of indomethacin. Monitoring of cyclooxygenase activity in an oxygraph also reveals an inhibition of PHS by BNF which depends only weakly on arachidonic acid concentration; inhibition by ANF is less pronounced under these conditions. Since PHS-catalyzed conversion of the benzoflavone compounds was detected under conditions permitting cooxidation, the inhibition of PHS by benzoflavones in vitro could either be a direct effect or possibly mediated via metabolites. Our data imply that ANF and BNF, in addition to their well-known role as modifiers of mixed-function oxidases, can affect the PHS-catalyzed metabolism of xenobiotics. This is discussed in the context of adverse effects caused by DES in vivo and in cell culture and must be taken into account when interpreting the modifying effect of benzoflavones on these endpoints.  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