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.
Adhesive properties of tenascin-X (TN-X) were investigated using TN-X purified from bovine skin and recombinant proteins encompassing the RGD sequence located within the tenth fibronectin type-III domain, and the fibrinogen-like domain. Osteosarcoma (MG63) and bladder carcinoma cells (ECV304) cells were shown to adhere to purified TN-X, but did not spread and did not assemble actin stress fibers. Both cell types adhered to recombinant proteins harboring the contiguous fibronectin type-III domains 9 and 10 (FNX 9-10) but not to the FNX 10 domain alone. This adhesion to FNX 9-10 was shown to be mediated by alphavbeta3 integrin, was inhibited by RGD peptides and was strongly reduced in proteins mutated within the RGD site. As antibodies against alphavbeta3 integrin had no effects on cell adhesion to purified TN-X, we suggest that the RGD sequence is masked in intact TN-X. Cell attachment to the recombinant TN-X fibrinogen domain (FbgX) and to purified TN-X was greater for MG63 than for ECV304 cells. A beta1-containing integrin was shown to be involved in MG63 cell attachment to FbgX and to purified TN-X. Although the existence of other cell interaction sites is likely in this huge molecule, these similar patterns of adhesion and inhibition suggest that the fibrinogen domain might be a dominant site in the whole molecule.  相似文献   

2.
Platelet endothelial cell adhesion molecule 1 (PECAM-1) (CD31), a member of the immunoglobulin (Ig) superfamily of cell adhesion molecules with six Ig-like domains, has a range of functions, notably its contributions to leukocyte extravasation during inflammation and in maintaining vascular endothelial integrity. Although PECAM-1 is known to mediate cell adhesion by homophilic binding via domain 1, a number of PECAM-1 heterophilic ligands have been proposed. Here, the possibility that heparin and heparan sulfate (HS) are ligands for PECAM-1 was reinvestigated. The extracellular domain of PECAM-1 was expressed first as a fusion protein with the Fc region of human IgG1 fused to domain 6 and second with an N-terminal Flag tag on domain 1 (Flag-PECAM-1). Both proteins bound heparin immobilized on a biosensor chip in surface plasmon resonance (SPR) binding experiments. Binding was pH-sensitive but is easily measured at slightly acidic pH. A series of PECAM-1 domain deletions, prepared in both expression systems, were tested for heparin binding. This revealed that the main heparin-binding site required both domains 2 and 3. Flag-PECAM-1 and a Flag protein containing domains 1-3 bound HS on melanoma cell surfaces, but a Flag protein containing domains 1-2 did not. Heparin oligosaccharides inhibited Flag-PECAM-1 from binding immobilized heparin, with certain structures having greater inhibitory activity than others. Molecular modeling similarly identified the junction of domains 2 and 3 as the heparin-binding site and further revealed the importance of the iduronic acid conformation for binding. PECAM-1 does bind heparin/HS but by a site that is distinct from that required for homophilic binding.  相似文献   

3.
We utilized recombinant fibronectin polypeptides with cell-binding domain and heparin-binding domains (referred to as C-274 and H-271, respectively) and their fusion polypeptide (CH-271) to examine the role of sulfated polysaccharide heparin and/or the functional domains of fibronectin in modulating tumor cell behavior. Both C-274 and CH-271 polypeptides with cell-binding domains promoted the adhesion and migration of B16-BL6 melanoma cells, whereas H-271 did not. Heparin bound to the immobilized polypeptides with heparin-binding domain (H-271, CH-271, and a mixture of C-274 and H-271 or fibronectin) but did not affect the tumor cell adhesion to the substrates. At the same time, heparin or two monoclonal antibodies against the heparin-binding domain were able to inhibit the haptotactic migration to CH-271 or fibronectin, though not to C-274 or a mixture of C-274 and H-271. This suggests that although heparin did not affect tumor cell adhesion to the cell-binding domain near the heparin-binding domain in CH-271 or fibronectin, it did lead to a modulation of cell motility. It seems likely that the regulatory mechanism may depend on interaction between heparin-like molecules on the cell surface and the heparin-binding domain in fibronectin, rather than on simple steric hindrance or on the masking of the cell-binding domain caused by the binding of heparin to heparin-binding domain.  相似文献   

4.
ColQ, the collagen tail subunit of asymmetric acetylcholinesterase, is responsible for anchoring the enzyme at the vertebrate synaptic basal lamina by interacting with heparan sulfate proteoglycans. To get insights about this function, the interaction of ColQ with heparin was analyzed. For this, heparin affinity chromatography of the complete oligomeric enzyme carrying different mutations in ColQ was performed. Results demonstrate that only the two predicted heparin-binding domains present in the collagen domain of ColQ are responsible for heparin interaction. Despite their similarity in basic charge distribution, each heparin-binding domain had different affinity for heparin. This difference is not solely determined by the number or nature of the basic residues conforming each site, but rather depends critically on local structural features of the triple helix, which can be influenced even by distant regions within ColQ. Thus, ColQ possesses two heparin-binding domains with different properties that may have non-redundant functions. We hypothesize that these binding sites coordinate acetylcholinesterase positioning within the organized architecture of the neuromuscular junction basal lamina.  相似文献   

5.
Interactions of heparin with intact human thrombospondin-1 (TSP1) and with two heparin-binding fragments of TSP1 were characterized using chemically modified heparins, a vascular heparan sulfate proteoglycan, and a series of heparin oligosaccharides prepared by partial deaminative cleavage. The avidity of TSP1 binding increased with oligosaccharide size, with plateaus at 4 to 6 and at 8 to 10 monosaccharide units. The dependence on oligosaccharide size for binding to the recombinant amino-terminal heparin-binding domain of TSP1 was the same as that of the intact TSP1 molecule but differed from that of a synthetic heparin-binding peptide from the type 1 repeats, suggesting that the interaction between intact TSP1 and heparin is primarily mediated by the amino-terminal domain. Based on activities of chemically modified heparins, binding to TSP1 depended primarily on 2-N- and 6-O-sulfation of glucosamine and to a lesser degree on 2,3-O-sulfation and the carboxyl residues of the uronic acids. In contrast, all of these modifications were required for binding of heparin to the type 1 repeat peptides. Affinity purification of heparin octasaccharides on immobilized TSP1 type 1 repeat peptides revealed a preference for oligosaccharides containing the disaccharide sequence IdoA(2-OSO(3))alpha1-4-GlcNS(6-OSO(3)). Binding of these oligosaccharides to the peptide required the Trp residues. These data demonstrate that the heparin-binding specificities of intact TSP1 and peptides from the type 1 repeats overlap with that of basic fibroblast growth factor (FGF2) and are consistent with the ability of these TSP1-derived molecules to inhibit FGF2-stimulated angiogenesis.  相似文献   

6.
Thrombospondin (TSP) is a homotrimeric extracellular glycoprotein with a subunit molecular mass of 140 kD. The subunits have a modular or domain-like structure and are held together by interchain disulphide bonds. A number of domains have been identified including those for the binding of collagen, fibrinogen, and heparin. Due to the trimeric form of the TSP molecule, the various domains are trivalent in nature and this contributes to the ability of TSP to mediate cell-substrate interactions. Indeed, TSP has recently been shown not only to promote cell adhesion but also to be intimately involved in cell growth and migration. The adhesive function of TSP is attributable to the "solid-phase" or matrix-bound form of the molecule. There is some evidence that the heparin-binding domain mediates incorporation of soluble TSP into the insoluble matrix form. The heparin-binding domain of TSP is a compact globular amino-terminal moiety that contains two clusters of basic amino acids and a single intrachain disulphide bond. To delineate the role of the heparin-binding domain in matrix assembly and to define further the precise region of interchain disulphide bonding that results in trimer formation, we have expressed deleted forms of the cDNA encoding TSP in SV-40-transformed. African green monkey kidney cells. The proteins synthesized from the various deleted TSP cDNAs were examined for (a) secretion into the culture medium and incorporation into the extracellular matrix; (b) binding to heparin-Sepharose; (c) immunoprecipitability by a conformation-specific monoclonal antibody; and (d) ability to form trimers. This analysis allowed us to draw the following conclusions. (a) A 218 amino acid NH2-terminal protein that preserves the intrachain disulphide bridge of the heparin-binding domain is capable of binding to heparin-Sepharose and incorporating into the extracellular matrix. (b) A shorter 164 amino acid NH2-terminal peptide that does not contain the intrachain disulphide bridge of the heparin-binding domain is neither able to bind to heparin-Sepharose nor able to incorporate into the extracellular matrix. (c) The region of interchain disulphide bridging necessary for trimer assembly resides within a cluster of seven cysteine residues immediately adjacent to the heparin-binding domain.  相似文献   

7.
Eleven structural analogues of human basic fibroblast growth factor (bFGF) have been prepared by site-directed mutagenesis of a synthetic bFGF gene to examine the effect of amino acid substitutions in the three putative heparin-binding domains on FGF's biological activity. After expression in Escherichia coli, the mutant proteins were purified to homogeneity by use of heparin-Sepharose chromatography and analyzed for their ability to stimulate DNA synthesis in human foreskin fibroblasts. Recombinant human bFGF 1-146 and [Ala69,Ser87]bFGF, an analogue where two of the four cysteines had been replaced by alanine and serine, were equipotent to standard bovine basic fibroblast growth factor. Substitution of aspartic acid-19 by arginine in the first heparin-binding domain yielded a molecule that stimulated a higher total mitogenic response in fibroblasts as compared to bFGF. In addition, replacement of either arginine-107 in the second domain or glutamine-123 in the third domain with glutamic acid resulted in compounds that were 2 and 4 times more potent than bFGF. In contrast, substitution of arginine-107 with isoleucine reduced the activity of the molecule by 100-fold. Combination of domain substitutions to generate the [Glu107,123]bFGF and [Arg19,Lys123,126]bFGF mutants did not show any additivity of the mutations on biological activity. Alterations in the biological activity of the analogues was dependent on both the site of and the type of modification. Increased positive charge in the first domain and increased negative charge in the second and third domains enhanced biological potency. The altered activities of the derivatives appear to be due in part to changes in the affinity of the analogues for heparin. We conclude that changes in all three of the putative heparin-binding domains result in altered mitogenic activity and heparin interaction of basic fibroblast growth factor.  相似文献   

8.
A primary heparin-binding site in vitronectin has been localized to a cluster of cationic residues near the C terminus of the protein. More recently, secondary binding sites have been proposed. In order to investigate whether the binding site originally identified on vitronectin functions as an exclusive and independent heparin-binding domain, solution binding methods have been used in combination with NMR and recombinant approaches to evaluate ligand binding to the primary site. Evaluation of the ionic strength dependence of heparin binding to vitronectin according to classical linkage theory indicates that a single ionic bond is prominent. It had been previously shown that chemical modification of vitronectin using an arginine-reactive probe results in a significant reduction in heparin binding (Gibson, A., Baburaj, K., Day, D. E., Verhamme, I. , Shore, J. D., and Peterson, C. B. (1997) J. Biol. Chem. 272, 5112-5121). The label has now been localized to arginine residues within the cyanogen bromide fragment-(341-380) that contains the primary heparin-binding site on vitronectin. One- and two-dimensional NMR on model peptides based on this primary heparin-binding site indicate that an arginine residue participates in the ionic interaction and that other nonionic interactions may be involved in forming a complex with heparin. A recombinant polypeptide corresponding to the C-terminal 129 amino acids of vitronectin exhibits heparin-binding affinity that is comparable to that of full-length vitronectin and is equally effective at neutralizing heparin anticoagulant activity. Results from this broad experimental approach argue that the behavior of the primary site is sufficient to account for the heparin binding activity of vitronectin and support an exposed orientation for the site in the structure of the native protein.  相似文献   

9.
The interaction between tenascin-C (TN-C), a multi-subunit extracellular matrix protein, and heparin was examined using a surface plasmon resonance-based technique on a Biacore system. The aims of the present study were to examine the affinity of fibronectin type III repeats of TN-C fragments (TNIII) for heparin, to investigate the role of the TNIII4 domains in the binding of TN-C to heparin, and to delineate a sequence of amino acids within the TNIII4 domain, which mediates cooperative heparin binding. At a physiological salt concentration, and pH 7.4, TNIII3-5 binds to heparin with high affinity (K(D) = 30 nm). However, a major heparin-binding site in TNIII5 produces a modest affinity binding at a K(D) near 4 microm, and a second site in TNIII4 enhances the binding by several orders of magnitude, although it was far too weak to produce an observable binding of TNIII4 by itself. Moreover, mutagenesis of the KEDK sequence in the TNIII4 domain resulted in the significant reduction of heparin-binding affinity. In addition, residues in the KEDK sequences are conserved in TN-C throughout mammalian evolution. Thus the structure-based sequence alignment, mutagenesis, and sequence conservation data together reveal a KEDK sequence in TNIII4 suggestive of a minor heparin-binding site. Finally, we demonstrate that TNIII4 contains binding sites for heparin sulfate proteoglycan and enhances the heparin sulfate proteoglycan-dependent human gingival fibroblast adhesion to TNIII5, thus providing the biological significance of heparin-binding site of TNIII4. These results suggest that the heparin-binding sites may traverse TNIII4-5 and thus require KEDK in TNIII4 for optimal heparin-binding.  相似文献   

10.
Secreted modular calcium-binding proteins 1 and 2 (SMOC-1 and SMOC-1) are extracellular calcium- binding proteins belonging to the BM-40 family of proteins. In this work we have identified a highly basic region in the extracellular calcium-binding (EC) domain of the SMOC-1 similar to other known glycosaminoglycan-binding motifs. Size-exclusion chromatography shows that full length SMOC-1 as well as its C-terminal EC domain alone bind heparin and heparan sulfate, but not the related chondroitin sulfate or dermatan sulfate glycosaminoglycans. Intrinsic tryptophan fluorescence measurements were used to quantify the binding of heparin to full length SMOC-1 and the EC domain alone. The calculated equilibrium dissociation constants were in the lower micromolar range. The binding site consists of two antiparallel alpha helices and mutagenesis experiments have shown that heparin-binding residues in both helices must be replaced in order to abolish heparin binding. Furthermore, we show that the SMOC-1 EC domain, like the SMOC-2 EC domain, supports the adhesion of epithelial HaCaT cells. Heparin-binding impaired mutants failed to support S1EC-mediated cell adhesion and together with the observation that S1EC in complex with soluble heparin attenuated cell adhesion we conclude that a functional and accessible S1EC heparin-binding site mediates adhesion of epithelial cells to SMOC-1.  相似文献   

11.
Solution structure of midkine, a new heparin-binding growth factor.   总被引:8,自引:0,他引:8       下载免费PDF全文
Midkine (MK) is a 13 kDa heparin-binding polypeptide which enhances neurite outgrowth, neuronal cell survival and plasminogen activator activity. MK is structurally divided into two domains, and most of the biological activities are located on the C-terminal domain. The solution structures of the two domains were determined by NMR. Both domains consist of three antiparallel beta-strands, but the C-terminal domain has a long flexible hairpin loop where a heparin-binding consensus sequence is located. Basic residues on the beta-sheet of the C-terminal domain form another heparin-binding site. Measurement of NMR signals in the presence of a heparin oligosaccharides verified that multiple amino acids in the two sites participated in heparin binding. The MK dimer has been shown to be the active form, giving signals to endothelial cells and probably to neuronal cells. We present a head-to-head dimer model of MK. The model was supported by the results of cross-linking experiments using transglutaminase. The dimer has a fused heparin-binding site at the dimer interface of the C-terminal domain, and the heparin-binding sites on MK fit the sulfate group clusters on heparin. These features are consistent with the proposed stronger heparin-binding activity and biological activity of the dimer.  相似文献   

12.
Abstract: Recent studies have identified the Alzheimer's disease amyloid β/A4 protein precursor (APP) as a trophic and/or tropic protein on several types of cells, including fibroblasts, primary culture neurons, PC12 cells, and B103 neuron-like cells. Many trophic proteins bind heparin, and it is believed that the heparin-binding domain is crucial for the trophic activity of these proteins. APP also binds heparin. The current studies were undertaken to examine the hypothesis that the neuritotropic activity of APP requires heparin binding. It was found that APP produced in E. coli bound B103 cells through detergent-extractable molecules. Approximately 50% of the binding sites were heparinase-sensitive, and heparin and heparan sulfate competed for APP binding to these sites. The heparinase-insensitive sites were recognized by a stretch of 17 amino acids of APP (residues 319–335) that contains the neuritotropic activity of APP. A mutant APP with a deletion at this site was capable of binding to the heparinase-sensitive sites, although this molecule was not neuritotropic to B103 neuron-like cells. Therefore, the neuritotropic site and the heparin-binding site are distinct in APP, and the neuritotropic effect of APP is produced through its binding to detergent-extractable and heparinase-insensitive sites.  相似文献   

13.
A minicollagen comprising the two C-terminal domains of collagen XII (COL1 and NC1) has been expressed in insect cells and characterized biochemically. An interaction with heparin is demonstrated, which depends on the correct folding of the molecule. After secretion, minicollagen XII is immediately processed to a form lacking heparin binding ability. Processed and unprocessed trimers differ only at the level of the eight or nine C-terminal residues but they reveal different structures as judged from rotary shadowing images. Similar processing is also observed in the medium of transfected human HeLa cells. These data show that a heparin-binding site is present in the C-terminal end of the chicken collagen XII sequence and strongly suggest that proteolytic processing in the NC1 domain can occur in vivo and regulate the interactive properties of collagen XII.  相似文献   

14.
Abstract: Recent studies have shown that the binding of the amyloid protein precursor (APP) of Alzheimer's disease to heparan sulfate proteoglycans (HSPGs) can modulate a neurite outgrowth-promoting function associated with APP. We used three different approaches to identify heparin-binding domains in APP. First, as heparin-binding domains are likely to be within highly folded regions of proteins, we analyzed the secondary structure of APP using several predictive algorithms. This analysis showed that two regions of APP695 contain a high degree of secondary structure, and clusters of basic residues, considered mandatory for heparin binding, were found principally within these regions. To determine which domains of APP bind heparin, deletion mutants of APP695 were prepared and analyzed for binding to a heparin affinity column. The results suggested that there must be at least two distinct heparin-binding regions in APP. To identify novel heparin-binding regions, peptides homologous to candidate heparin-binding domains were analyzed for their ability to bind heparin. These experiments suggested that APP contains at least four heparin-binding domains. The presence of more than one heparin-binding domain on APP suggests the possibility that APP may interact with more than one type of glycosaminoglycan.  相似文献   

15.
Localization of the major heparin-binding site in fibronectin   总被引:15,自引:0,他引:15  
We have identified the major site required for the interaction of fibronectin (FN) with heparin. Affinity chromatography was used to test the binding ability of a library of truncated, monomeric forms of fibronectin (deminectins) containing deletions or two point mutations in the heparin-binding domain. This domain consists of type III repeats 12, 13, and 14. Deletions of individual repeats showed that both III13 and III14 are required for complete binding. Small deletions within these repeats localized a major site of heparin interaction to the amino-terminal half of III13. Site-directed mutagenesis of adjacent arginines within this sequence to uncharged residues reduced heparin binding by 98%, identifying these positively charged amino acids as essential for the interaction. A significant role for the flanking alternatively spliced regions and for repeat III12 was not found. We conclude that, while both repeats III13 and III14 participate in heparin binding, there is a major site of interaction in repeat III13 that accounts for nearly all of the activity. The significance of multiple heparin-binding sites within this domain is discussed and a model is proposed to account for how these sites may function in vivo.  相似文献   

16.
Chen FF  Lin WH  Lin SC  Kuo JH  Chu HY  Huang WC  Chuang YJ  Lee SC  Sue SC 《Glycobiology》2012,22(5):649-661
Hepatoma-derived growth factor (HDGF) recognizes cell surface heparan sulfate to promote its internalization though binding to its N-terminal HATH (homologous to amino terminus of HDGF) domain. HDGF-related proteins (HRPs) all have the HATH domain in their N terminus. In this study, we report on the commonality of heparin binding in all HRPs with a broad range of heparin-binding affinity: HRP-4 is the strongest binder, and the lens epithelium-derived growth factor shows a relatively weak binding, with binding affinities (K(D)) showing 30-fold difference in magnitude. With the HDGF HATH domain used as a model, residue K19 was the most critical basic residue in molecular recognition and protein internalization, and with its proximal proline-tryptophan-tryptophan-proline motif, coordinated a conformational change when binding to the heparin fragment. Other basic residues, K21, K61, K70, K72 and R79, confer added contribution in binding that the total ionic interaction from these residues represents more than 70% of the binding energy. Because the positive-charged residues are conserved in all HRP HATH domains, heparin binding outside of cells might be of equal importance for all HRPs in mediating downstream signaling; however, distinct effects and/or distribution might be associated with the varying affinities to heparin.  相似文献   

17.
The red cell ICAM-4/LW blood group glycoprotein, which belongs to the family of intercellular adhesion molecules (ICAMs), has been reported to interact with CD11a/CD18 (LFA-1) and CD11b/CD18 (Mac-1) beta(2) integrins. To better define the basis of the ICAM-4/beta(2) integrin interaction, we have generated wild-type, domain-deleted and mutated recombinant chimeric ICAM-4-Fc proteins and analyzed their interaction in a cellular adhesion assay with LFA-1 and Mac-1 L-cell stable transfectants. We found that monoclonal antibodies against CD11a, CD11b, CD18, or LW(ab) block adhesion of transfectant L-cells to immobilized ICAM-4-Fc protein and that the ICAM-4/beta(2) integrin interaction was highly sensitive to the presence of the divalent cations Ca(2+) and Mg(2+). Deletion of individual Ig-domains D1 or D2 of the extracellular part of ICAM-4 showed that LFA-1 binds to the first Ig-like domain, whereas the Mac-1 binding site encompassed both the first and the second Ig-like domains. Based on the crystal structure of ICAM-2, we propose a model for the Ig-like domains D1 and D2 of ICAM-4. Accordingly, by site-directed mutagenesis of 22 amino acid positions spread out on all faces of the ICAM-4 molecule, we identified four exposed residues, Leu(80), Trp(93), and Arg(97) on the CFG face and Trp(77) on the E-F loop of domain D1 that may contact LFA-1 as part of the binding site. However, the single and double mutants R52E and T91Q on the CFG face of domain D1, which correspond to the key residues Glu(34) and Gln(73) for ICAM-1 binding to LFA-1, had no effect on LFA-1 binding. In contrast, all mutants on the CFG face of domain D1 and residues Glu(151) and Thr(154) in the C'-E loop of the domain D2 seem to play a dominant role in Mac-1 binding. These data suggest that the binding site for LFA-1 on ICAM-4 overlaps but is distinct from the Mac-1 binding site.  相似文献   

18.
Yamazaki Y  Tokunaga Y  Takani K  Morita T 《Biochemistry》2005,44(24):8858-8864
VEGF-A165 displays multiple effects through binding to KDR (VEGFR-2). Heparin/heparan sulfate-like molecules are known to greatly modulate their interaction. In fact, VEGF-A lacking a C-terminal heparin-binding region exhibits significantly reduced mitogenic activity. We recently found novel heparin-binding VEGFs in snake venom, designated VEGF-Fs, which specifically recognize KDR, rather than other VEGF receptors. VEGF-Fs virtually lack the C-terminal heparin-binding region when compared with other heparin-binding VEGF subtypes, despite their heparin-binding potential. The C-terminal region does not exhibit any significant homology with other known proteins or domains. In this study, we attempted to identify the heparin-binding region of VEGF-F using synthetic peptides. The C-terminal peptide of vammin (one of the VEGF-Fs, 19 residues) bound to heparin with similar affinity as native vammin. We then evaluated the effects of this peptide on the biological activity of VEGF-A165. The C-terminal peptide of VEGF-F exhibited specific blockage of VEGF-A165 activity both in vitro and in vivo. These observations demonstrate that the short C-terminal region of VEGF-F functions fully as the active heparin-binding domain and the corresponding peptide specifically blocks VEGF-A165, thus suggesting that the C-terminal heparin-binding region of VEGF-F recognizes similar heparin/heparan sulfate molecules as VEGF-A165. The present results will provide novel insight into VEGF-heparin interaction and may facilitate the design of new anti-VEGF drugs based on novel strategies.  相似文献   

19.
Annexin A2 and heparin bind to one another with high affinity and in a calcium-dependent manner, an interaction that may play a role in mediating fibrinolysis. In this study, three heparin-derived oligosaccharides of different lengths were co-crystallized with annexin A2 to elucidate the structural basis of the interaction. Crystal structures were obtained at high resolution for uncomplexed annexin A2 and three complexes of heparin oligosaccharides bound to annexin A2. The common heparin-binding site is situated at the convex face of domain IV of annexin A2. At this site, annexin A2 binds up to five sugar residues from the nonreducing end of the oligosaccharide. Unlike most heparin-binding consensus patterns, heparin binding at this site does not rely on arrays of basic residues; instead, main-chain and side-chain nitrogen atoms and two calcium ions play important roles in the binding. Especially significant is a novel calcium-binding site that forms upon heparin binding. Two sugar residues of the heparin derivatives provide oxygen ligands for this calcium ion. Comparison of all four structures shows that heparin binding does not elicit a significant conformational change in annexin A2. Finally, surface plasmon resonance measurements were made for binding interactions between annexin A2 and heparin polysaccharide in solution at pH 7.4 or 5.0. The combined data provide a clear basis for the calcium dependence of heparin binding to annexin A2.  相似文献   

20.
Fibroblast growth factors (FGFs) interact with heparan sulfate glycosaminoglycans and the extracellular domains of FGF cell surface receptors (FGFRs) to trigger receptor activation and biological responses. FGF homologous factors (FHF1-FHF4; also known as FGF11-FGF14) are related to FGFs by substantial sequence homology, yet their only documented interactions are with an intracellular kinase scaffold protein, islet brain-2 (IB2) and with voltage-gated sodium channels. In this report, we show that recombinant FHFs can bind heparin with high affinity like classical FGFs yet fail to activate any of the seven principal FGFRs. Instead, we demonstrate that FHFs bind IB2 directly, furthering the contention that FHFs and FGFs elicit their biological effects by binding to different protein partners. To understand the molecular basis for this differential target binding specificity, we elucidated the crystal structure of FHF1b to 1.7-A resolution. The FHF1b core domain assumes a beta-trefoil fold consisting of 12 antiparallel beta strands (beta 1 through beta 12). The FHF1b beta-trefoil core is remarkably similar to that of classical FGFs and exhibits an FGF-characteristic heparin-binding surface as attested to by the number of bound sulfate ions. Using molecular modeling and structure-based mutational analysis, we identified two surface residues, Arg52 in the beta 4-beta 5 loop and Val95 in the beta 9 strand of FHF1b that are required for the interaction of FHF1b with IB2. These two residues are unique to FHFs, and mutations of the corresponding residues of FGF1 to Arg and Val diminish the capacity of FGF1 to activate FGFRs, suggesting that these two FHF residues contribute to the inability of FHFs to activate FGFRs. Hence, FHFs and FGFs bear striking structural similarity but have diverged to direct related surfaces toward interaction with distinct protein targets.  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