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.
The serine proteinase inhibitor antithrombin III (ATIII) is a key regulatory protein of intrinsic blood coagulation. ATIII attains its full biological activity only upon binding polysulfated oligosaccharides, such as heparin. A series of synthetic peptides have been prepared based on the proposed heparin binding regions of ATIII and their ability to bind heparin has been assessed by CD spectrometry, by isothermal titration calorimetry, and by the ability of the peptides to compete with ATIII for binding heparin in a factor Xa procoagulant enzyme assay. Peptide F123-G148, which encompasses both the purported high-affinity pentasaccharide binding region and an adjacent, C-terminally directed segment of ATIII, was found to bind heparin with good affinity, but amino-terminal truncations of this sequence, including L130-G148 and K136-G148 displayed attenuated heparin binding activities. In fact, K136-G148 appears to encompass only a low-affinity heparin binding site. In contrast, peptides based solely on the high-affinity binding site (K121-A134) displayed much higher affinities for heparin. By CD spectrometry, these high-affinity peptides are chiefly random coil in nature, but low microM concentrations of heparin induce significant alpha-helix conformation. K121-A134 also effectively competes with ATIII for binding heparin. Thus, through the use of synthetic peptides that encompass part, if not all, of the heparin binding site(s) within ATIII, we have further elucidated the structure-function relations of heparin-ATIII interactions.  相似文献   

2.
A heparin-binding peptide within antithrombin III (ATIII) was identified by digestion of ATIII with Staphylococcus aureus V8 protease followed by purification on reverse-phase high pressure liquid chromatography using a C-4 column matrix. The column fractions were assayed for their ability to bind heparin by ligand blotting with 125I-fluoresceinamine-heparin as previously described (Smith, J. W., and Knauer, D. J. (1987) Anal. Biochem. 160, 105-114). This analysis identified at least three fractions with heparin binding ability of which the peptide eluting at 25.4 min gave the strongest signal. Amino acid sequence analysis of this peptide gave a partially split sequence which was consistent with regions encompassing amino acids 89-96 and 114-156. These amino acids are present in a 1:1 molar ratio which is consistent with a disulfide linkage between Cys-95 and Cys-128. High affinity heparin competed more effectively for the binding of 125I-fluoresceinamine-heparin to this peptide than low affinity heparin. Chondroitin sulfate did not block the binding of 125I-fluoresceinamine-heparin to the peptide. These data strongly suggest that the isolated peptide represents a native heparin-binding region within intact ATIII. Computer generation of a plot of running charge density of ATIII confirms that the region encompassing amino acid residues 123-141 has the highest positive charge density within the molecule. A hydropathy plot of ATIII was generated using a method similar to that of Kyte and Doolittle (Kyte, J., and Doolittle, R. F. (1982) J. Mol. Biol. 157, 105-132). This plot indicates that amino acid residues 126-140 are exposed to the exterior surface of the molecule. Based on these data, we suggest that the region corresponding to amino acid residues 114-156 is a likely site for the physiological heparin-binding domain of ATIII. We also conclude that the proposed disulfide bridges within the protein are suspect and should be re-examined (Petersen, T. E., Dudek-Wojiechowska, G., Sottrup-Jensen, L., and Magnussun, S. (1979) in The Physiological Inhibitors of Coagulation and Fibrinolysis (Collen, D., Wiman, B., and Verstaeta, M., eds) pp. 43-54, Elsevier Scientific Publishing Co., Amsterdam).  相似文献   

3.
Nonrandom structural features in the heparin polymer   总被引:1,自引:0,他引:1  
Computer simulation studies were used to prepare an ensemble of heparin number chains. The polydispersity of these chains was simulated by introducing a specific "fraction of terminators", and it closely resembled the experimentally observed polydispersity of a porcine mucosal, glycosaminoglycan heparin. The same percentage of simulated chains contained antithrombin III (ATIII) binding site sequences as are typically found to contain ATIII binding sites using affinity chromatography. Heparin lyase action was then simulated by using Michaelis-Menten kinetics. In one model, heparin chains were constructed from the random assembly of monosaccharide units using the observed mole percentage of each. After simulated depolymerization, the final oligosaccharides formed were compared to the observed oligosaccharide products. The simulation which assumed a random distribution of monosaccharide units in heparin did not agree with experimental observations. In particular, no ATIII binding site sequences were found in the simulated number chains. The results of this simulation indicate that heparin is not simply a random assembly of monosaccharide units. These results are consistent with the known, ordered biosynthesis of heparin. In a second model, heparin chains were constructed from randomly assembled oligosaccharides at the mole percentage in which each is found in the final product mixture. The action of heparin lyase was then simulated, and the distribution of the oligosaccharide products was measured throughout the simulated time course of the depolymerization reaction. The simulated rate of formation and final concentration of a particular oligosaccharide which contains a portion of heparin's ATIII binding site were similar to those observed experimentally. These results are consistent with the random distribution of ATIII binding sites within glycosaminoglycan heparin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
In the blood coagulation cascade, human antithrombin III (hAT III) acts as an inhibitor of serine proteases such as thrombin and factor Xa, and this anticoagulatory glycoprotein requires the binding of heparin for its activation. In this study, we synthesized the polypeptides corresponding to the proposed heparin-binding sites including the (41-49), (286-301) and (123-139) regions of hAT III, and examined their interactions with heparin by means of physicochemical and biochemical methods. All the synthetic peptides had a high affinity toward heparin, evidenced by the fact that they were eluted from a heparin-agarose column at the high salt concentration range of 520-700 mM. In addition, hAT III (123-139) attenuated the effect of heparin on the activation of hAT III, whereas other HBPs did not, suggesting that only hAT III (123-139) could interact with the active site of heparin. On the basis of these results, we prepared novel hAT III (123-139)-related derivatives as potent heparin antagonist candidates, and examined the influence of several modifications on their activity in vitro. The results provided new findings about the structure-activity relationship of hAT III (123-139), and led us to the successful development of a potent antagonist for heparin.  相似文献   

5.
The last step of heparin biosynthesis is thought to involve the action of 3-O-sulfotransferase resulting in the formation of an antithrombin III (ATIII) binding site required for heparin's anticoagulant activity. The isolation of a significant fraction of heparin chains without antithrombin III-binding sites and having low affinity for ATIII suggests the presence of a precursor site, lacking the 3-O-sulfate group. Porcine mucosal heparin was depolymerized into a mixture of oligosaccharides using heparin lyase. One of these oligosaccharides was derived from heparin's ATIII-binding site. In an effort to find the ATIII-binding site precursor, the structures of several minor oligosaccharides were determined. A greater than 90% recovery of oligosaccharides (on a mole and weight basis) was obtained for both unfractionated and affinity-fractionated heparins. An oligosaccharide arising from the ATIII-binding site precursor was found that comprised only 0.8 mol % of the oligosaccharide product mixture. This oligosaccharide was only slightly enriched in heparin having a low affinity for ATIII and only slightly disenriched in high affinity heparin. The small number of these ATIII-binding site precursors, found in unfractionated and fractionated heparins, suggests the existence of a low ATIII affinity heparin may not simply be the result of the incomplete action of 3-O-sulfotransferase in the final step in heparin biosynthesis. Rather these data suggest that some earlier step, involved in the formation of placement of these precursor sites, may be primarily responsible for high and low ATIII affinity heparins.  相似文献   

6.
The binding of heparin to human antithrombin III (ATIII) was investigated by titration calorimetry (TC) and differential scanning calorimetry (DSC). TC measurements of homogeneous high-affinity pentasaccharide and octasaccharide fragments of heparin in 0.02 M phosphate buffer and 0.15 M sodium chloride (pH 7.3) yielded binding constants of (7.1 +/- 1.3) x 10(5) M-1 and (6.7 +/- 1.2) x 10(6) M-1, respectively, and corresponding binding enthalpies of -48.3 +/- 0.7 and -54.4 +/- 5.4 kJ mol-1. The binding enthalpy of heparin in phosphate buffer (0.02 M, 0.15 M NaCl, pH 7.3) was estimated from TC measurements to be -55 +/- 10 kJ mol-1, while the enthalpy in Tris buffer (0.02 M, 0.15 M NaCl, pH 7.3) was -18 +/- 2 kJ mol-1. The heparin-binding affinity was shown by fluorescence measurements not to change under these conditions. The 3-fold lower binding enthalpy in Tris can be attributed to the transfer of a proton from the buffer to the heparin-ATIII complex. DSC measurements of the ATIII unfolding transition exhibited a sharp denaturation peak at 329 +/- 1 K with a van 't Hoff enthalpy of 951 +/- 89 kJ mol-1, based on a two-state transition model and a much broader transition from 333 to 366 K. The transition peak at 329 K accounted for 9-18% of the total ATIII. At sub-saturate heparin concentrations, the lower temperature peak became bimodal with the appearance of a second transition peak at 336 K. At saturate heparin concentration only the 336 K peak was observed. This supports a two domain model of ATIII folding in which the lower stability domain (329 K) binds and is stabilized by heparin.  相似文献   

7.
An hereditary abnormal antithrombin III (ATIII Geneva) with defective heparin cofactor activity was characterized by DNA single strand amplification and subsequent direct sequencing. ATIII Geneva was found to have a G to A transition in Exon IIIa leading to an Arg-129 to Gln mutation. This amino acid is part of the ATIII region comprising residues 114-154, which contains the highest proportion of basic residues (Arg or Lys), and is known from chemical modification studies to be involved in heparin binding. The variant protein did not bind heparin-Sepharose and was isolated from the propositus plasma by immunoaffinity chromatography. High affinity (for ATIII) heparin had only a minimal effect on thrombin and activated factor X inhibition by the purified abnormal ATIII. Taken together, these results demonstrate an important role for Arg-129 in the binding and interaction of ATIII with heparin of high affinity. We propose that a cooperation between Lys-125, Arg-129, Lys-136, and Arg-47 exposed at the surface of the inhibitor allows the binding of the essential pentasaccharide domain of heparin which is specific for the ATIII interaction.  相似文献   

8.
Antithrombin III (ATIII) is the main inhibitor of the coagulation proteases like factor Xa and thrombin. Anticoagulant activity of ATIII is increased by several thousand folds when activated by vascular wall heparan sulfate proteoglycans (HSPGs) and pharmaceutical heparins. ATIII isoforms in human plasma, alpha-ATIII and beta-ATIII differ in the amount of glycosylation which is the basis of differences in their heparin binding affinity and function. Crystal structures and site directed mutagenesis studies have mapped the heparin binding site in ATIII, however the hydrogen bond switch and energetics of interaction during the course of heparin dependent conformational change remains largely unclear. An analysis of heparin bound conformational states of ATIII using PEARLS software showed that in heparin bound intermediate state, Arg 47 and Arg 13 residues make hydrogen bonds with heparin but in the activated conformation Lys 11 and Lys 114 have more hydrogen bond interactions. In the protease bound-antithrombin-pentasaccharide complex Lys 114, Pro 12 and Lys 125 form important hydrogen bonding interactions. The results showed that A-helix and N-terminal end residues are more important in the initial interactions but D-helix is more important during the latter stage of conformational activation and during the process of protease inhibition. We carried out the residue wise Accessible Surface Area (ASA) analysis of alpha and beta ATIII native states and the results indicated major differences in burial of residues from Ser 112 to Ser 116 towards the N-terminal end. This region is involved in the P-helix formation on account of heparin binding. A cavity analysis showed a progressively larger cavity formation during activation in the region just adjacent to the heparin binding site towards the C-terminal end. We hypothesize that during the process of conformational change after heparin binding beta form of antithrombin has low energy barrier to form D-helix extension toward N and C-terminal end as compared to alpha isoform.  相似文献   

9.
Porcine mucosal heparin was fragmented into low-molecular-weight (LMW) heparin by treatment of periodate-oxidized heparin with sodium hydroxide, followed by reduction with sodium borohydride and acid hydrolysis. Gradient polyacrylamide gel electrophoresis analysis showed a mixture of heparin fragments with an average size of eight disaccharide units. 1D 1H NMR showed two-thirds of the N-acetyl groups were lost on periodate cleavage, suggesting cleavage had occurred at the glucopyranosyluronic acid (GlcpA) and idopyranosyluronic acid (IdopA) residues located within and adjacent to the antithrombin III (ATIII) binding site. The N-acetyl glucopyranose (GlcpNAc) residue was lost on workup. The GlcpA residue, within the ATIII binding site, is on the non-reducing side of the N-sulfo, 3, 6-O-sulfo glycopyranosylamine (GlcpNS3S6S) residue. Thus, periodate cleaved heparin should be enriched in GlcpNS3S6S residues. Two-dimensional correlation spectroscopy (2D COSY) confirmed that LMW heparin prepared through periodate cleavage contained GlcpNS3S6S at its non-reducing end. As expected, this LMW heparin also showed reduced ATIII mediated anti-factor IIa and anti-factor Xa activities.  相似文献   

10.
Antithrombin III (ATIII) is a key antiproteinase involved in blood coagulation. Previous investigations have shown that ATIII is degraded by Staphylococcus aureus V8 protease, leading to release of heparin binding fragments derived from its D helix. As heparin binding and antimicrobial activity of peptides frequently overlap, we here set out to explore possible antibacterial effects of intact and degraded ATIII. In contrast to intact ATIII, the results showed that extensive degradation of the molecule yielded fragments with antimicrobial activity. Correspondingly, the heparin-binding, helix d-derived, peptide FFFAKLNCRLYRKANKSSKLV (FFF21) of human ATIII, was found to be antimicrobial against particularly the Gram-negative bacteria Escherichia coli and Pseudomonas aeruginosa. Fluorescence microscopy and electron microscopy studies demonstrated that FFF21 binds to and permeabilizes bacterial membranes. Analogously, FFF21 was found to induce membrane leakage of model anionic liposomes. In vivo, FFF21 significantly reduced P. aeruginosa infection in mice. Additionally, FFF21 displayed anti-endotoxic effects in vitro. Taken together, our results suggest novel roles for ATIII-derived peptide fragments in host defense.  相似文献   

11.
Three sulphated polysaccharides, dermatan sulphate, fucan and heparin, were fractionated according to their affinity towards antithrombin III (ATIII) and heparin cofactor II (HCII), the two main physiological thrombin (IIa) inhibitors. Both inhibitors were immobilized on concanavalin A—Sepharose, which binds to the glycosylated chains of the proteins while the protein-binding site for the polysaccharide remains free. Each polysaccharide was fractionated into bound and unbound fractions either for ATIII or HCII. The eluted fractions were tested for their ability to catalyse and interactions. The possible presence of a unique binding site for ATIII and HCII, on each sulphated polysaccharide, was also studied.  相似文献   

12.
Only 30% of commercial heparin reacts with antithrombin III (ATIII). This study shows that the interaction is pH dependent: 100% of the heparin binds to ATIII at pH 3.0, 30% at physiological pH. Binding of ATIII, platelet factor 4, and protamine to heparin was studied using a new rocket precipitin electrophoresis procedure, adapted from the Laurell rocket immunoelectrophoresis procedure. Protamine is incorporated into agarose gel, and heparin mixtures with protamine, ATIII, or platelet factor 4 electrophoresed into the gel from a series of wells. The residual free heparin is precipitated by the protamine in a rocket-shaped arc, the height of which is proportional to the amount of free heparin. No antibody is employed. This procedure is useful for quantitation of heparin and for studying the binding of heparin to proteins.  相似文献   

13.
Arocas V  Turk B  Bock SC  Olson ST  Björk I 《Biochemistry》2000,39(29):8512-8518
The interaction of a well-defined pentasaccharide sequence of heparin with a specific binding site on antithrombin activates the inhibitor through a conformational change. This change increases the rate of antithrombin inhibition of factor Xa, whereas acceleration of thrombin inhibition requires binding of both inhibitor and proteinase to the same heparin chain. An extended heparin binding site of antithrombin outside the specific pentasaccharide site has been proposed to account for the higher affinity of the inhibitor for full-length heparin chains by interacting with saccharides adjacent to the pentasaccharide sequence. To resolve conflicting evidence regarding the roles of Lys136 and Lys139 in this extended site, we have mutated the two residues to Ala or Gln. Mutation of Lys136 decreased the antithrombin affinity for full-length heparin by at least 5-fold but minimally altered the affinity for the pentasaccharide. As a result, the full-length heparin and pentasaccharide affinities were comparable. The reduced affinity for full-length heparin was associated with the loss of one ionic interaction and was caused by both a lower overall association rate constant and a higher overall dissociation rate constant. In contrast, mutation of Lys139 affected neither full-length heparin nor pentasaccharide affinity. The rate constants for inhibition of thrombin and factor Xa by the complexes between antithrombin and full-length heparin or pentasaccharide were unaffected by both mutations, indicating that neither Lys136 nor Lys139 is involved in heparin activation of the inhibitor. Together, these results show that Lys136 forms part of the extended heparin binding site of antithrombin that participates in the binding of full-length heparin chains, whereas Lys139 is located outside this site.  相似文献   

14.
In the blood coagulation cascade, heparin activates human plasma antithrombin III (hAT III), resulting in the inhibition of factor Xa. This polysaccharide also exhibits hemorrhagic tendency mediated by the inhibition of thrombin in heparinotherapy. Therefore, attention has focused on the development of low molecular weight heparins (LMW-heparins) that inhibit factor Xa but not thrombin. In this investigation, we examined the biochemical and physicochemical properties of hAT III-derived heparin-binding peptides (HBPs). Of all the tested HBPs, hAT III (123-139) exhibited the highest affinity with heparin and showed an inhibitory effect on the heparin-induced enhancement of hAT III activity toward factor Xa, indicating that hAT III (123-139) specifically interacts with the active region in heparin. We prepared a synthetic hAT III (123-139)-coupled affinity chromatography system, and demonstrated that this novel affinity chromatography is useful for fractionation of highly active moieties in LMW-heparins.  相似文献   

15.
Heparin cofactor II (Mr = 65,600) was purified 1800-fold from human plasma to further characterize the structural and functional properties of the protein as they compare to antithrombin III (Mr = 56,600). Heparin cofactor II and antithrombin III are functionally similar in that both proteins have been shown to inhibit thrombin at accelerated rates in the presence of heparin. There was little evidence for structural homology between heparin cofactor II and antithrombin III when high performance liquid chromatography-tryptic peptide maps and NH2-terminal sequences were compared. A partially degraded form of heparin cofactor II was also obtained in which a significant portion (Mr = 8,000) of the NH2 terminus was missing. The rates of thrombin inhibition (+/- heparin) by native and partially degraded-heparin cofactor II were not significantly different, suggesting that the NH2-terminal region of the protein is not essential either for heparin binding or for thrombin inhibition. A significant degree of similarity was found in the COOH-terminal regions of the proteins when the primary structures of the reactive site peptides, i.e. the peptides which are COOH-terminal to the reactive site peptide bonds cleaved by thrombin, were compared. Of the 36 residues identified, 19 residues in the reactive site peptide sequence of heparin cofactor II could be aligned with residues in the reactive site peptide from antithrombin III. While the similarities in primary structure suggest that heparin cofactor II may be an additional member of the superfamily of proteins consisting of antithrombin III, alpha 1-antitrypsin, alpha 1-antichymotrypsin and ovalbumin, the differences in structure could account for differences in protease specificity and reactivity toward thrombin. In particular, a disulfide bond which links the COOH-terminal (reactive site) region of antithrombin III to the remainder of the molecule and is important for the heparin-induced conformational change in the protein and high affinity binding of heparin does not appear to exist in heparin cofactor II. This observation provides an initial indication that while the reported kinetic mechanisms of action of heparin in accelerating the heparin cofactor II/thrombin and antithrombin III/thrombin reactions are similar, the mechanisms and effects of heparin binding to the two inhibitors may be different.  相似文献   

16.
The dissociation equilibrium constant for heparin binding to antithrombin III (ATIII) is a measure of the cofactor's binding to and activation of the proteinase inhibitor, and its salt dependence indicates that ionic and non-ionic interactions contribute approximately 40 and approximately 60% of the binding free energy, respectively. We now report that phenylalanines 121 and 122 (Phe-121 and Phe-122) together contribute 43% of the total binding free energy and 77% of the energy of non-ionic binding interactions. The large contribution of these hydrophobic residues to the binding energy is mediated not by direct interactions with heparin, but indirectly, through contacts between their phenyl rings and the non-polar stems of positively charged heparin binding residues, whose terminal amino and guanidinium groups are thereby organized to form extensive and specific ionic and non-ionic contacts with the pentasaccharide. Investigation of the kinetics of heparin binding demonstrated that Phe-122 is critical for promoting a normal rate of conformational change and stabilizing AT*H, the high affinity-activated binary complex. Kinetic and structural considerations suggest that Phe-122 and Lys-114 act cooperatively through non-ionic interactions to promote P-helix formation and ATIII binding to the pentasaccharide. In summary, although hydrophobic residues Phe-122 and Phe-121 make minimal contact with the pentasaccharide, they play a critical role in heparin binding and activation of antithrombin by coordinating the P-helix-mediated conformational change and organizing an extensive network of ionic and non-ionic interactions between positively charged heparin binding site residues and the cofactor.  相似文献   

17.
Since the discovery of anti-HIV activity in oligo(tyrosine sulfate)s in our laboratory, we have been interested in their potential as heparin pentasaccharide mimics. In this study, we investigated their interactions with synthetic heparin-binding peptides, derived from human antithrombin III (hAT III) and heparin-interacting protein (HIP), using surface noncovalent affinity mass spectrometry. We compared binding affinities to those heparin-binding peptides between oligo(tyrosine sulfate)s and several known sulfated compounds and found that oligo(tyrosine sulfate)s bind to hAT III (123-139) more strongly than a heparin-derived hexasaccharide dp6. Moreover, we found longer oligo(tyrosine sulfate) has higher binding affinity to hAT III (123-139).  相似文献   

18.
The specific localization of the asymmetric form of acetylcholinesterase (AChE) in neuromuscular junctions results from the interaction of its collagen-like tail with heparan sulfate proteoglycans in the synaptic basal lamina. This interaction involves two heparin binding consensus sequences of the form XBBXB, where B is a basic residue, located in the triple-helical collagen tail: GRKGR for the N-terminal site and GKRGK for the C-terminal site. To explore the basis of the higher heparin affinity seen for the C-terminal site vs. the N-terminal site, two homologous series of (Gly-Xaa-Yaa)(8) peptides were constructed to model these triple-helical binding sites. Individual tripeptide units from each heparin binding site were introduced in a stepwise fashion into a Gly-Pro-Hyp framework, until the consensus sequence and its surrounding triplets were recreated. As each additional triplet from the binding site is inserted to replace a host Gly-Pro-Hyp triplet, the triple-helix stability decreases, and the drop in thermal stability is close to that expected if each Gly-X-Y triplet contributed independently to global stability. CD spectroscopy and calorimetry show the stability of these AChE model peptides is increased by addition of heparin, confirming binding to heparin, and the lack of significant enthalpy change indicates the binding is largely electrostatic in nature. Displacement assays measure the strength of the peptide-heparin interaction, and indicate an inverse correlation between the peptide ability to bind heparin and its thermal stability. The model peptides for the C-terminal binding site show a greater heparin affinity than the peptide models for the N-terminal binding site only when residues surrounding the consensus sequence are included.  相似文献   

19.
A dysfunctional antithrombin III (ATIII) gene encoding a qualitatively and quantitatively abnormal anticoagulant molecule is responsible for hereditary thrombosis in a Utah kindred [Bock et al. (1985) Am. J. Hum. Genet. 37, 32-41]. Nucleotide sequencing of the entire protein-encoding portion of the cloned ATIII-Utah gene revealed a C to T transitional mutation which converts proline-407 to leucine. Proline-407 is located 14 amino acids C-terminal to the reactive site arginine of ATIII in a core region of the molecule that has been highly conserved during evolution of the serine protease inhibitor (serpin) gene family. The location of this proline in the crystal structure of the homologous serpin alpha 1-antitrypsin suggests that the leucine substitution in ATIII-Utah may interfere with correct folding of the mutant gene product, leading to its rapid turnover and the low antithrombin levels observed in patient plasmas. The Pro-407 to Leu mutation does not interfere with binding of antithrombin III to heparin. Patient antithrombin III, isolated by affinity chromatography on heparin-Sepharose, was reacted with purified thrombin. ATIII encoded by the patient's normal gene formed protease-inhibitor complexes with thrombin, whereas the product of the ATIII-Utah gene did not. The Pro-407 to Leu mutation destroys a restriction site for the enzyme StuI, permitting rapid diagnosis of affected members of the Utah kindred by Southern blotting of genomic DNA.  相似文献   

20.
CD36 (glycoprotein IV or IIIB) is an integral plasma membrane protein of wide cellular distribution and functions as a receptor site for thrombospondin (TSP), an adhesive protein important in cell-cell and cell-matrix interactions. OKM5, a monoclonal anti-CD36 antibody, has been reported to block CD36 cell adhesive functions suggesting that the OKM5 epitope on CD36 is functionally important. A panel of 10 synthetic CD36 peptides was made. One peptide, P139-155, specifically inhibited the immunoadsorption of CD36 by OKM5, and P139-155 was directly immunoadsorbed by OKM5, indicating that CD36 sequence 139-155 represents part of the OKM5 epitope. TSP bound to immobilized P139-155 in a dose-dependent and saturable manner. Surprisingly, P139-155 significantly augmented, instead of inhibited, binding of CD36 to TSP. This peptide did not induce platelet aggregation but augmented ADP- and collagen-induced aggregation in platelet-rich plasma. Another CD36 peptide, P93-110, which had no effect on OKM5 immunoadsorption, blocked binding of CD36 to immobilized TSP and partially inhibited collagen-induced platelet aggregation. P93-110 by itself did not bind to TSP; however, in the presence of P139-155, there was a marked enhancement of P93-110 binding to TSP, with a stoichiometry consistent with the trimeric nature of TSP. The data suggest that CD36-TSP interaction is a two-step process; the sequence 139-155 region of CD36 binds first to TSP, triggering a change in TSP to reveal a second site, which binds the 93-110 region of CD36 with high affinity. CD36 peptides can be used as stimulators or inhibitors in cellular adhesive events involving TSP-CD36 interaction. Conformational changes leading to the exposure or activation of high affinity binding sites may occur in both the receptor and the ligand upon cell-cell and cell-matrix adhesion.  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