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.
Mitsi M  Hong Z  Costello CE  Nugent MA 《Biochemistry》2006,45(34):10319-10328
Regulation of angiogenesis involves interactions between vascular endothelial growth factor (VEGF) and components of the extracellular matrix, including fibronectin and heparan sulfate. In the present study, we identified two classes of VEGF binding sites on fibronectin. One was constitutively available whereas the availability of the other was modulated by the conformational state of fibronectin. Atomic force microscopy studies revealed that heparin and hydrophilic substrates promoted the extended conformation of fibronectin, leading to increased VEGF binding. The ability of heparin to enhance VEGF binding to fibronectin was dependent on the chemical composition and chain length of heparin, since long (>22 saccharides) heparin chains with sulfation on the 6-O and N positions of glucosamine units were required for full activity. Treatment of the complex endothelial extracellular matrix with heparin also increased VEGF binding, suggesting that heparin/heparan sulfate might regulate VEGF interactions within the extracellular matrix by controlling the structure and organization of fibronectin matrices.  相似文献   

2.
Hypoxia is one of the major signals that induces angiogenesis. Hypoxic conditions lead to reduced extracellular pH. Vascular endothelial growth factor (VEGF) binding to endothelial cells and the extracellular matrix (ECM) increases at acidic pH (7.0-5.5). These interactions are dependent on heparan sulfate proteoglycans, but do not depend on the presence of VEGF receptors. Here we report that VEGF(165) and VEGF(121) binding to fibronectin also increased at acidic pH, and that these interactions are further enhanced by the addition of heparin. These results reveal that the accepted non-heparin-binding isoform of VEGF (VEGF(121)) is converted into a heparin-binding growth factor under acidic conditions. Interestingly, we did not observe increased binding of VEGF to collagen type I at acidic pH in the presence or absence of heparin, indicating that this effect is not a general property of all heparin-binding ECM proteins. The high level of VEGF binding at acidic pH was also rapidly reversed as demonstrated by increased rates of VEGF dissociation from fibronectin and fibronectin-heparin matrices as the pH was raised. The VEGF released from fibronectin retained its ability to stimulate the activation of extracellular-regulated kinase 1/2 in endothelial cells. These results suggest that VEGF may be stored in the extracellular matrix via interactions with fibronectin and heparan sulfate in tissues that are in need of vascularization so that it can aid in directing the dynamic process of growth and migration of new blood vessels.  相似文献   

3.
Thrombospondin is a major glycoprotein of the platelet alpha-granule and is secreted during platelet activation. Several protease-resistant domains of thrombospondin mediate its interactions with components of the extracellular matrix including fibronectin, collagen, heparin, laminin, and fibrinogen. Thrombospondin, as well as fibronectin, is composed of several discretely located biologically active domains. We have characterized the thrombospondin binding domains of plasma fibronectin and determined the binding affinities of the purified domains; fibronectin has at least two binding sites for thrombospondin. Thrombospondin bound specifically to the 29-kDa amino-terminal heparin binding domain of fibronectin as well as to the 31-kDa non-heparin binding domain located within the larger 40-kDa carboxy-terminal fibronectin domain generated by chymotrypsin proteolysis. Platelet thrombospondin interacted with plasma fibronectin in a specific and saturable manner in blot binding as well as solid-phase binding assays. These interactions were independent of divalent cations. Thrombospondin bound to the 29-kDa fibronectin heparin binding domain with a Kd of 1.35 x 10(-9) M. The Kd for the 31-kDa domain of fibronectin was 2.28 x 10(-8) M. The 40-kDa carboxy-terminal fragment bound with a Kd of 1.65 x 10(-8) M. Heparin, which binds to both proteins, inhibited thrombospondin binding to the amino-terminal domain of fibronectin by more than 70%. The heparin effect was less pronounced with the non-heparin binding carboxy-terminal domain of fibronectin. By contrast, the binding affinity of the thrombospondin 150-kDa domain, which itself lacked heparin binding, was not affected by the presence of heparin. Based on these data, we conclude that thrombospondin binds with different affinities to two distinct domains in the fibronectin molecule.  相似文献   

4.
Streptococcal fibronectin binding protein I (SfbI) mediates adherence to and invasion of Streptococcus pyogenes into human epithelial cells. In this study, we analysed the binding activity of distinct domains of SfbI protein towards its ligand, the extracellular matrix component fibronectin, as well as the biological implication of the binding events during the infection process. By using purified recombinant SfbI derivatives as well as in vivo expressed SfbI domains on the surface of heterologous organism Streptococcus gordonii , we were able to dissociate the two major streptococcal target domains on the human fibronectin molecule. The SfbI repeat region exclusively bound to the 30 kDa N-terminal fragment of fibronectin, whereas the SfbI spacer region exclusively bound to the 45 kDa collagen-binding fragment of fibronectin. In the case of native surface-expressed SfbI protein, an induced fit mode of bacteria–fibronectin interaction was identified. We demonstrate that binding of the 30 kDa fibronectin fragment to the repeat region of SfbI protein co-operatively activates the adjacent SfbI spacer domain to bind the 45 kDa fibronectin fragment. The biological consequence arising from this novel mode of fibronectin targeting was analysed in eukaryotic cell invasion assays. The repeat region of SfbI protein is mediating adherence and constitutes a prerequisite for subsequent invasion, whereas the SfbI spacer domain efficiently triggers the invasion process of streptococci into the eukaryotic cell. Thus, we were able to dissect bacterial adhesion from invasion by manipulating one protein. SfbI protein therefore represents a highly evolved prokaryotic molecule that exploits the host factor fibronectin not only for extracellular targeting but also for its subsequent activation that leads to efficient cellular invasion.  相似文献   

5.
Vascular endothelial growth factor A (VEGF-A) is a promoter of neovascularization and thus a popular therapeutic target for diseases involving excessive growth of blood vessels. In this study, we explored the potential of the disaccharide sucrose octasulfate (SOS) to alter VEGF165 diffusion through Descemet's membrane. Descemet's membranes were isolated from bovine eyes and used as a barrier between two chambers of a diffusion apparatus to measure VEGF transport. Diffusion studies revealed a dramatic increase in VEGF165 transport in the presence of SOS, with little diffusion of VEGF165 across the membrane over a 10-h time course in the absence of SOS. Diffusion studies with VEGF121, a non-heparin binding variant of VEGF, showed robust diffusion with or without SOS. To determine a possible mechanism, we measured the ability of SOS to inhibit VEGF interactions with extracellular matrix (ECM), using cell-free and cell surface binding assays. Binding studies showed SOS had no effect on VEGF165 binding to either heparin-coated plates or endothelial cell surfaces at less than mg/ml concentrations. In contrast, we show that SOS inhibited VEGF165 binding to fibronectin in a dose dependent manner and dramatically accelerated the rate of release of VEGF165 from fibronectin. SOS also inhibited the binding of VEGF165 to fibronectin-rich ECM deposited by vascular smooth muscle cells. These results suggest that fibronectin-rich extracellular matrices serve as barriers to VEGF165 diffusion by providing a network of binding sites that can trap and sequester the protein. Since the content of Descemet's membrane is typical of many basement membranes it is possible that they serve throughout the body as formidable barriers to VEGF165 diffusion and tightly regulate its bioavailability and distribution within tissues.  相似文献   

6.
Introduction of Salmonella enterica serotype Typhimurium into food products results from its ability to persist in the intestine of healthy livestock by mechanisms that are poorly understood. The non-fimbrial adhesin ShdA is a fibronectin binding protein required for persistent intestinal carriage of S. Typhimurium. We further investigated the molecular mechanism of ShdA-mediated intestinal persistence by determining the binding-site of this receptor in fibronectin. Analysis of ShdA binding to fibronectin proteolytic fragments and to recombinant fibronectin fusion proteins identified the (13)FnIII repeat module of the Hep-2 domain as the primary binding site for this adhesin. The (13)FnIII repeat module of fibronectin contains a cationic cradle formed by six basic residues (R6, R7, R9, R23, K25 and R54) that is a high affinity heparin-binding site conserved among fibronectin sequences from frogs to man. Binding of ShdA to the (13)FnIII repeat module of fibronectin and to a second extracellular matrix protein, Collagen I, could be inhibited by heparin. Furthermore, binding of ShdA to the Hep-2 domain was sensitive to the ionic buffer strength, suggesting that binding involved ionic interactions. We therefore determined whether amino acid substitutions of basic residues in the cationic cradle of the Hep-2 domain that inhibit heparin binding also abrogate binding of ShdA. Combined substitution of R6S and R7S strongly reduced ShdA binding to (13)FnIII. These data suggest that ShdA binds the Hep-2 domain of fibronectin by a mechanism that may mimic binding of the host polysaccharide heparin.  相似文献   

7.
We examined the mechanisms of cell attachment to fibronectin-coated substrates. Inhibition of cell attachment was obtained by species- specific antifibronectin antibodies, which presumably recognize a distinct antigenic structure in the protein located at, or in the immediate vicinity of, the cell-binding site. The inhibiting antibodies could be adsorbed on a column of Sepharose substituted with plasma fibronectin. The initial phase of cell attachment was also inhibited by addition of soluble fibronectin to the incubation medium in a reaction that exhibited specificity and concentration dependence. These data suggest that cell-binding sites are available in an active form on the surface of soluble fibronectin. However, the inhibitory effect of fibronectin was greatly enhanced by adding the protein together with heparin, heparan sulfate, collagen, or a fibronectin-binding collagen peptide (CB-7), which is consistent with an "activation" of fibronectin on binding to these matrix components. A similar activation of fibronectin was obtained by cleaving the protein with trypsin. We discuss these findings in relation to conformational rearrangements in the fibronectin molecule. Data is presented supporting a mechanism of cell attachment to fibronectin involving multiple weak interactions between cellular receptors and substrate molecules, although some steps in the attachment process appear to disobey the requirements for such a mechanism.  相似文献   

8.
Thrombospondin is a large multifunctional glycoprotein synthesized, secreted and incorporated into the extracellular matrix by several cell types in culture. It is also present in the blood platelet and is secreted following platelet activation. We have previously shown that thrombospondin co-distributes with fibronectin in the extracellular matrix and that it can bind directly to purified fibronectin. In order to elucidate the chemical aspects of thrombospondin incorporation into the extracellular matrix, we studied the interaction of endothelial cell thrombospondin and fibronectin. We find that endothelial cell thrombospondin has two distinct binding domains for fibronectin. One domain is on the 70-kDa core fragment, probably similar to that of platelet thrombospondin. The other domain is on the 27-kDa N-terminal fragment and is unique to endothelial cell thrombospondin. The dissociation constant of the intact endothelial-cell-derived molecule is 0.7 +/- 0.2 x 10(-7) M. Following fragmentation, the separate domains bind with somewhat lower affinity: the core domain binds with a Kd of 3.4 +/- 1.5 x 10(-7) M and the N-terminal domain binds with a Kd of 8.8 +/- 1.8 x 10(-7) M. Binding of the intact molecule is Ca2+-independent. By contrast, following tryptic fragmentation, binding of the 70-kDa fragment is practically lost. It can be restored, however, by removal of Ca2+, indicating that the binding site on this domain is either sequestered or becomes so following fragmentation. Heparin, which also binds to both fragments, competed with fibronectin binding to the 27-kDa fragment but not to the 70-kDa domain. The fact that heparin also competitively inhibits fibronectin binding of the intact molecule further supports sequestration of the fibronectin-binding domain on the 70-kDa core fragment. Our data suggest that endothelial-cell thrombospondin possesses two distinct binding sites for fibronectin, a low-affinity constitutively available one and a high-affinity one, possibly sequestered on the intact unbound molecule.  相似文献   

9.
Transforming growth factor beta (TGF-beta) enhances the cell surface binding of 125I-fibronectin by cultured human fibroblasts. The effect of TGF-beta on cell surface binding was maximal after 2 h of exposure to TFG-beta and did not require epidermal growth factor or protein synthesis. The enhancement was dose dependent and was found with the 125I-labeled 70-kilodalton amino-terminal fragment of fibronectin as well as with 125I-fibronectin. Treatment of cultures with TGF-beta for 6 h resulted in a threefold increase in the estimated number of fibronectin binding sites. The increase in number of binding sites was accompanied by an increased accumulation of labeled fibronectin in detergent-insoluble extracellular matrix. The effect of TGF-beta was biphasic; after 6 h of exposure, less labeled fibronectin bound to treated cultures than to control cultures. Exposure of cells to TGF-beta for greater than 6 h caused a two- to threefold increase in the accumulation of cellular fibronectin in culture medium as detected by a quantitative enzyme-linked immunosorbent assay. The second phase of the biphasic effect and the increase in soluble cellular fibronectin were blocked by cycloheximide. Immunofluorescence staining of fibroblast cultures with antifibronectin revealed that TGF-beta caused a striking increase in fibronectin fibrils. The 70-kilodalton amino-terminal fragment of fibronectin, which blocks incorporation of fibronectin into extracellular matrix, blocked anchorage-independent growth of NRK-49F cells in the presence of epidermal growth factor. Our results show that an increase in the binding and rate of assembly of exogenous fibronectin is an early event preceding the increase in expression of extracellular matrix proteins. Such an early increase in cell surface binding of exogenous fibronectin may be a mechanism whereby TGF-beta can modify extracellular matrix characteristics rapidly after tissue injury or during embryonic morphogenesis.  相似文献   

10.
The lymphocyte-high endothelial venule (HEV) cell interaction is an essential element of the immune system, as it controls lymphocyte recirculation between blood and lymphoid organs in the body. This interaction involves an 85-95-kD class of lymphocyte surface glycoprotein(s), CD44. A subset of lymphocyte CD44 molecules is modified by covalent linkage to chondroitin sulfate (Jalkanen, S., M. Jalkanen, R. Bargatze, M. Tammi, and E. C. Butcher. 1988. J. Immunol. 141:1615-1623). In this work, we show that removal of chondroitin sulfate by chondroitinase treatment of lymphocytes or incubation of HEV with chondroitin sulfate does not significantly inhibit lymphocyte binding to HEV, suggesting that chondroitin sulfate is not involved in endothelial cell recognition of lymphocytes. Affinity-purified CD44 antigen was, on the other hand, observed to bind native Type I collagen fibrils, laminin, and fibronectin, but not gelatin. Binding to fibronectin was studied more closely, and it was found to be mediated through the chondroitin sulfate-containing form of the molecule. The binding site on fibronectin was the COOH-terminal heparin binding domain, because (a) the COOH-terminal heparin-binding fragment of fibronectin-bound isolated CD44 antigen; (b) chondroitin sulfate inhibited this binding; and (c) finally, the ectodomain of another cell surface proteoglycan, syndecan, which is known to bind the COOH-terminal heparin binding domain of fibronectin (Saunders, S., and M. Bernfield. 1988. J. Cell Biol. 106: 423-430), inhibited binding of CD44 both to intact fibronectin and to its heparin binding domain. Moreover, inhibition studies showed that binding of a lymphoblastoid cell line, KCA, to heparin binding peptides from COOH-terminal heparin binding fragment of fibronectin was mediated via CD44. These findings suggest that recirculating lymphocytes use the CD44 class of molecules not only for binding to HEV at the site of lymphocyte entry to lymphoid organs as reported earlier but also within the lymphatic tissue where CD44, especially the subset modified by chondroitin sulfate, is used for interaction with extracellular matrix molecules such as fibronectin.  相似文献   

11.
Latent transforming growth factor-beta-binding proteins (LTBPs) are extracellular matrix (ECM) glycoproteins that play a major role in storage of latent TGF-beta in the ECM and regulate its availability. We have previously identified fibronectin as a key molecule for incorporation of LTBP1 and TGF-beta into the ECM of osteoblasts and fibroblasts. Here we provide evidence that heparan sulfate proteoglycans may mediate binding between LTBP1 and fibronectin. We have localized critical domains in the N terminus of LTBP1 that are required for co-localization with fibronectin in osteoblast cultures and have identified heparin binding sites in the N terminus of LTBP1 between residues 345 and 487. Solid-phase binding assays suggest that LTBP1 does not bind directly to fibronectin but that the binding is indirect. Heparin coupled to bovine serum albumin (heparin-BSA) was able to mediate binding between fibronectin and LTBP1. Treatment of primary osteoblast cultures with heparin or heparin-BSA but not with chondroitin sulfate impaired LTBP1 deposition onto fibronectin without inhibiting expression of LTBP1. Inhibition of LTBP1 incorporation was accompanied by reduced incorporation of latent TGF-beta into the ECM, with increased amounts of soluble latent TGF-beta. Inhibition of attachment of glycosaminoglycans to the core proteins of proteoglycans by beta-d-xylosides also reduced incorporation of LTBP1 into the ECM. These studies suggest that heparan sulfate proteoglycans may play a critical role in regulating TGF-beta availability by controlling the deposition of LTBP1 into the ECM in association with fibronectin.  相似文献   

12.
The longer splice isoforms of vascular endothelial growth factor-A (VEGF-A), including mouse VEGF164, contain a highly basic heparin-binding domain (HBD), which imparts the ability of these isoforms to be deposited in the heparan sulfate-rich extracellular matrix and to interact with the prototype sulfated glycosaminoglycan, heparin. The shortest isoform, VEGF120, lacks this highly basic domain and is freely diffusible upon secretion. Although the HBD has been attributed significant relevance to VEGF-A biology, the molecular determinants of the heparin-binding site are unknown. We used site-directed mutagenesis to identify amino acid residues that are critical for heparin binding activity of the VEGF164 HBD. We focused on basic residues and found Arg-13, Arg-14, and Arg-49 to be critical for heparin binding and interaction with extracellular matrix in tissue samples. We also examined the cellular and biochemical consequences of abolishing heparin-binding function, measuring the ability of the mutants to interact with VEGF receptors, induce endothelial cell gene expression, and trigger microvessel outgrowth. Induction of tissue factor expression, vessel outgrowth, and binding to VEGFR2 were unaffected by the HBD mutations. In contrast, the HBD mutants showed slightly decreased binding to the NRP1 (neuropilin-1) receptor, and analyses suggested the heparin and NRP1 binding sites to be distinct but overlapping. Finally, mutations that affect the heparin binding activity also led to an unexpected reduction in the affinity of VEGF164 binding specifically to VEGFR1. This finding provides a potential basis for previous observations suggesting enhanced potency of VEGF164 versus VEGF120 in VEGFR1-mediated signaling in inflammatory cells.  相似文献   

13.
Angiogenesis, the growth of new blood vessels, is regulated by a number of factors, including hypoxia and vascular endothelial growth factor (VEGF). Although the effects of hypoxia have been studied intensely, less attention has been given to other extracellular parameters such as pH. Thus, the present study investigates the consequences of acidic pH on VEGF binding and activity in endothelial cell cultures. We found that the binding of VEGF165 and VEGF121 to endothelial cells increased as the extracellular pH was decreased from 7.5 to 5.5. Binding of VEGF165 and VEGF121 to endothelial extracellular matrix was also increased at acidic pH. These effects were, in part, a reflection of increased heparin binding, because VEGF165 and VEGF121 showed increased retention on heparin-Sepharose at pH 5.5 compared with pH 7.5. Consistent with these findings, soluble heparin competed for VEGF binding to endothelial cells under acidic conditions. However, at neutral pH (7.5) low concentrations of heparin (0.1-1.0 microg/ml) potentiated VEGF binding. Extracellular pH also regulated VEGF activation of the extracellular signal-regulated kinases 1 and 2 (Erk1/2). VEGF165 and VEGF121 activation of Erk1/2 at pH 7.5 peaked after 5 min, whereas at pH 6.5 the peak was shifted to 10 min. At pH 5.5, neither VEGF isoform was able to activate Erk1/2, suggesting that the increased VEGF bound to the cells at low pH was sequestered in a stored state. Therefore, extracellular pH might play an important role in regulating VEGF interactions with cells and the extracellular matrix, which can modulate VEGF activity.  相似文献   

14.
Heparin and HS (heparan sulfate) exert their wide range of biological activities by interacting with extracellular protein ligands. Among these important protein ligands are various angiogenic growth factors and cytokines. HS binding to VEGF (vascular endothelial growth factor) regulates multiple aspects of vascular development and function through its specific interaction with HS. Many studies have focused on HS-derived or HS-mimicking structures for the characterization of VEGF165 interaction with HS. Using a heparinase 1-prepared small library of heparin-derived oligosaccharides ranging from hexasaccharide to octadecasaccharide, we systematically investigated the heparin-specific structural features required for VEGF binding. We report the apparent affinities for the association between the heparin-derived oligosaccharides with both VEGF165 and VEGF55, a peptide construct encompassing exclusively the heparin-binding domain of VEGF165. An octasaccharide was the minimum size of oligosaccharide within the library to efficiently bind to both forms of VEGF and a tetradecasaccharide displayed an effective binding affinity to VEGF165 comparable to unfractionated heparin. The range of relative apparent binding affinities among VEGF and the panel of heparin-derived oligosaccharides demonstrate that the VEGF binding affinity likely depends on the specific structural features of these oligosaccharides, including their degree of sulfation, sugar-ring stereochemistry and conformation. Notably, the unique 3-O-sulfo group found within the specific antithrombin binding site of heparin is not required for VEGF165 binding. These findings afford new insight into the inherent kinetics and affinities for VEGF association with heparin and heparin-derived oligosaccharides with key residue-specific modifications and may potentially benefit the future design of oligosaccharide-based anti-angiogenesis drugs.  相似文献   

15.
16.
Tumor cell adhesion to the extracellular matrix is an important consideration in tumor metastasis. Recent results show that multiple adhesion-promoting domains for melanoma cells can be purified from proteolytic digests of fibronectin [McCarthy, J. B., Hagen, S. T., & Furcht, L. T. (1986) J. Cell Biol. 102, 179-188]. Monoclonal antibodies were generated against a tryptic/catheptic 33K heparin binding fragment of fibronectin derived from the carboxyl terminal of the A chain. This region contains a tumor cell adhesion-promoting domain(s). The amino-terminal sequence was determined for this fragment, as well as a tryptic 31K fragment which is located to the carboxyl-terminal side of the 33K heparin binding fragment in A chains of fibronectin. The partial sequence data demonstrate that arginyl-glycyl-aspartyl-serine (RGDS) or the related arginyl-glutamyl-aspartyl-valine (REDV) is not present in the 33K heparin binding fragment, confirming earlier results which demonstrated that cells adhere to this fragment by an RGDS-independent mechanism. Two monoclonal antibodies, termed AHB-1 and AHB-2, recognized epitopes common to heparin binding fragments derived from the carboxyl terminus of both the A and B chains of fibronectin. Monoclonal antibody AHB-2 inhibited melanoma adhesion to the 33K heparin binding fragment of fibronectin in a concentration-dependent manner, whereas monoclonal antibody AHB-1 had no effect on adhesion to this fragment. Neither monoclonal antibody inhibited adhesion to intact fibronectin. However, monoclonal AHB-2 potentiated the inhibitory effect of suboptimal levels of exogenous RGDS on cell adhesion to intact fibronectin. AHB-2 recognized an epitope common to both the A- and B-chain carboxyl-terminal heparin binding region of fibronectin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The extracellular matrix of cultured human lung fibroblasts contains one major heparan sulfate proteoglycan. This proteoglycan contains a 400-kDa core protein and is structurally and immunochemically identical or closely related to the heparan sulfate proteoglycans that occur in basement membranes. Because heparitinase does not release the core protein from the matrix of cultured cells, we investigated the binding interactions of this heparan sulfate proteoglycan with other components of the fibroblast extracellular matrix. Both the intact proteoglycan and the heparitinase-resistant core protein were found to bind to fibronectin. The binding of 125I-labeled core protein to immobilized fibronectin was inhibited by soluble fibronectin and by soluble cold core protein but not by albumin or gelatin. A Scatchard plot indicates a Kd of about 2 x 10(-9) M. Binding of the core protein was also inhibited by high concentrations of heparin, heparan sulfate, or chrondroitin sulfate and was sensitive to high salt concentrations. Thermolysin fragmentation of the 125I-labeled proteoglycan yielded glycosamino-glycan-free core protein fragments of approximately 110 and 62 kDa which bound to both fibronectin and heparin columns. The core protein-binding capacity of fibronectin was very sensitive to proteolysis. Analysis of thermolytic and alpha-chymotryptic fragments of fibronectin showed binding of the intact proteoglycan and of its isolated core protein to a protease-sensitive fragment of 56 kDa which carried the gelatin-binding domain of fibronectin and to a protease-sensitive heparin-binding fragment of 140 kDa. Based on the NH2-terminal amino acid sequence analyses of the 56- and 140-kDa fragments, the core protein-binding domain in fibronectin was tentatively mapped in the area of overlap of the two fragments, carboxyl-terminally from the gelatin-binding domain, possibly in the second type III repeat of fibronectin. These data document a specific and high affinity interaction between fibronectin and the core protein of the matrix heparan sulfate proteoglycan which may anchor the proteoglycan in the matrix.  相似文献   

18.
Analysis of parameters governing heparin binding to fibronectin indicates that heparin binding is a necessary, but insufficient, condition for fibronectin cryoprecipitation. Heparin binding to fibronectin is a rapid, readily reversible event which can occur under several conditions which prohibit fibronectin cryoprecipitation. While cryoprecipitation of fibronectin is abolished at temperatures in excess of 10 degrees C, appreciable heparin binding to fibronectin does occur even at 40 degrees C. While increasing ionic strength and pH inhibit both heparin binding and cryoprecipitation of fibronectin, heparin binding can still occur at high ionic strengths and pH values which completely abolish cryoprecipitation. Scatchard analysis of fluorescent polarization data reveals a biphasic heparin binding curve with high and low affinity Kd values of 3.5 X 10(-8) and 10(-6) M, respectively. In contrast to heparin binding, fibronectin aggregation is a cooperative phenomenon. Fibronectin cryoprecipitation is greatly reduced at temperatures above 10 degrees C, at pH values above pH 10, and at ionic strengths above 0.3 M. Thus, heparin binding and protein aggregation are separate events which occur during fibronectin cryoprecipitation. Results obtained here via fluorescence polarization in conjunction with other physical measurements suggest that a decrease in flexibility of the fibronectin molecule is associated with the protein aggregation step of cryoprecipitation. The role of heparin in the mechanism of fibronectin cryoprecipitation is discussed.  相似文献   

19.
Studies with cultured fibroblasts have shown that plasma as well as cellular fibronectin can be organized into fibrillar structures and that this organization is mediated by sites at the cell surface. Treatment of human skin fibroblasts with cholera toxin resulted in a prompt decrease in the number of binding sites for 125I-labeled plasma fibronectin and a 125I-labeled 70-kDa amino-terminal fragment of fibronectin. This decrease was accompanied by less incorporation of labeled fibronectin into deoxycholate-insoluble extracellular matrix. Binding of 125I-fibronectin was also decreased in cultures treated with epinephrine, isoproterenol, or forskolin. These results, therefore, indicate that G proteins and the adenylate cyclase system are involved in regulation of fibronectin matrix assembly sites may be one mechanism whereby hormones or growth factors can modify extracellular matrix characteristics.  相似文献   

20.
In the present study, we show that intact Mycoplasma fermentans cells have a wealth of adhesive interactions with components of the extracellular matrix. Mycoplasma fermentans intensively bind plasminogen, and to a lesser extent, fibronectin, heparin, and laminin. The binding of collagen type III, IV, or V was low. The binding of plasminogen, collagen type III, or collagen type V markedly enhanced the adherence of M. fermentans to HeLa cells, whereas the binding of fibronectin, heparin, laminin, or collagen IV induced only a small effect on mycoplasma adherence. Utilizing plasminogen-treated M. fermentans preparations, we detected microorganisms within host HeLa cells by the gentamicin protection assay or by confocal laser scanning microscopy of immunofluorescent preparations. However, no intracellular M. fermentans was detected when M. fermentans preparations treated with fibronectin, heparin, laminin, or collagen type III, IV, or V were utilized.  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