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.
The interaction of heparin with glia-derived nexin (GDN) has been characterized and compared to that observed between heparin and antithrombin III (ATIII). Heparin was fractionated according to its affinity for immobilized GDN, and the ability of various fractions to accelerate the inhibition rate of thrombin by either GDN or ATIII was examined. Fractions with different affinities for GDN accelerated the thrombin-GDN reaction to a similar extent; heparin with a high affinity for immobilized GDN stimulated the reaction only about 30% more than the fraction that did not bind to immobilized GDN. Slightly greater differences were observed for the effect of these fractions on the thrombin-ATIII reaction; heparin that did not bind to the GDN affinity column was about 60% more effective than heparin with a high affinity for GDN in accelerating the inhibition of thrombin by ATIII. The CNBr fragment of GDN between residues 63 and 144 was able to reduce the heparin-accelerated rate of inhibition of thrombin by GDN indicating that this region of GDN was able to bind the heparin molecules responsible for the acceleration. Shorter synthetic peptides within this sequence did not significantly reduce the rate, suggesting that the heparin-binding activity of fragment 63-144 depends on a specific conformation of the polypeptide chain. Fragment 63-144 was less effective in decreasing the heparin-accelerated rate of inhibition of thrombin by ATIII. The results are discussed in terms of the heparin species that are responsible for the acceleration of the GDN- and ATIII-thrombin reactions and the heparin-binding sites of GDN and ATIII.  相似文献   

2.
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   

3.
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   

4.
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   

5.
Harrop  HA; Rider  CC 《Glycobiology》1998,8(2):131-137
We have employed a direct radiolabel binding assay to investigate the interaction between3H-heparin and recombinant envelope glycoproteins, rgp120s, derived from several different isolates of HIV-1. Comparable dose-dependent binding is exhibited by rgp120s from isolates IIIB, GB8, MN and SF-2. Under identical experimental conditions the binding of3H- heparin to a recombinant soluble form of the cellular receptor for gp120, CD4, is negligible. The binding of3H-heparin to rgp120 is competed for by excess unlabeled heparin and certain other, but not all, glycosaminoglycan and chemically modified heparins. Of a range of such polysaccharides tested, ability to compete with3H-heparin for binding was strictly correlated with inhibition of HIV-1 replication in vitro. Those possessing potent anti-HIV-1 activity were effective competitors, whereas those having no or little anti-HIV-1 activity were poor competitors. Scatchard analysis indicates that the K d of the interaction between heparin and rgp120 is 10 nM. Binding studies conducted in increasing salt concentrations confirm that the interaction is ionic in nature. Synthetic 33-35 amino acid peptides based on the sequence of the V3 loop of gp120 also bind to heparin with high affinity. V3 loop peptides that are cyclized due to terminal cysteine residues show more selective binding than their uncyclized counterparts. Overall, these data demonstrate further that heparin exerts its anti-HIV-1 activity by binding to the envelope glycoprotein of HIV-1, rather than its cellular receptor, CD4. This study confirms that the V3 loop of gp120 is the site at which heparin exerts its anti- HIV-1 activity. Moreover, it reveals that high affinity binding to heparin is shared by all four rgp120s examined, despite amino acid substitutions within the V3 loop.   相似文献   

6.
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   

7.
Exploring the development of nonviral nucleic acid delivery vectors with progressive, specific, and novel designs in molecular architecture is a fundamental way to investigate how aspects of chemical and physical structure impact the transfection process. In this study, macromolecules comprised of a four-arm star poly(ethylene glycol) and termini modified with one of five different heparin binding peptides have been investigated for their ability to bind, compact, and deliver DNA to mammalian cells in vitro. These new delivery vectors combine a PEG-derived stabilizing moiety with peptides that exhibit unique cell-surface binding ability in a molecular architecture that permits multivalent presentation of the cationic peptides. Five peptide sequences of varying heparin binding affinity were studied; each was found to sufficiently bind heparin for biological application. Additionally, the macromolecules were able to bind and compact DNA into particles of proper size for endocytosis. In biological studies, the PEG-star peptides displayed a range of toxicity and transfection efficiency dependent on the peptide identity. The vectors equipped with peptides of highest heparin binding affinity were found to bind DNA tightly, increase levels of cellular internalization, and display the most promising transfection qualities. Our results suggest heparin binding peptides with specific sequences hold more potential than nonspecific cationic polymers to optimize transfection efficiency while maintaining cell viability. Furthermore, the built-in multivalency of these macromolecules may allow simultaneous binding of both DNA at the core of the polyplex and heparan sulfate on the surface of the cell. This scheme may facilitate a bridging transport mechanism, tethering DNA to the surface of the cell and subsequently ushering therapeutic nucleic acids into the cell. This multivalent star shape is therefore a promising architectural feature that may be exploited in the design of future polycationic gene delivery vectors.  相似文献   

8.
Transglutaminase 2 (TG2) is an autoantigen in celiac disease (CD) and it has multiple biologic functions including involvement in cell adhesion through interactions with integrins, fibronectin (FN), and heparan sulfate proteoglycans. We aimed to delineate the heparin‐binding regions of human TG2 by studying binding kinetics of the predicted heparin‐binding peptides using surface plasmon resonance method. In addition, we characterized immunogenicity of the TG2 peptides and their effect on cell adhesion. The high‐affinity binding of human TG2 to the immobilized heparin was observed, and two TG2 peptides, P1 (amino acids 202–215) and P2 (261–274), were found to bind heparin. The amino acid sequences corresponding to the heparin‐binding peptides were located close to each other on the surface of the TG2 molecule as part of the α‐helical structures. The heparin‐binding peptides displayed increased immunoreactivity against serum IgA of CD patients compared with other TG2 peptides. The cell adhesion reducing effect of the peptide P2 was revealed in Caco‐2 intestinal epithelial cell attachment to the FN and FN‐TG2 coated surfaces. We propose that TG2 amino acid sequences 202–215 and 261–274 could be involved in binding of TG2 to cell surface heparan sulfates. High immunoreactivity of the corresponding heparin‐binding peptides of TG2 with CD patient's IgA supports the previously described role of anti‐TG2 autoantibodies interfering with this interaction. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
A peptide model for the heparin binding site of antithrombin III (ATIII) was synthesized to elucidate the structural consequences of heparin binding. This peptide [ATIII(123-139)] and a sequence-permuted analogue (ATIII random) showed similar conformational behavior (as analyzed by circular dichroism spectroscopy) in aqueous and organic media. In the presence of heparin, however, the peptide ATIII(123-139) assumed a stable conformation, whereas peptide ATIII random did not. Complex formation was saturable and sensitive to salt. The ATIII(123-139)-heparin complex contained beta-structure, rather than helical structure. This finding is incompatible with current models of heparin binding and suggests that heparin binding may induce nonnative structures at the binding site which could, in turn, lead to activation of ATIII. The peptide ATIII(123-139) was able to inhibit the binding of ATIII by heparin, consistent with the notion that this peptide may be a model for the heparin binding site.  相似文献   

10.
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   

11.
The kinetics and energetics of the binding of three troponin-I peptides, corresponding to regions 96-131 (TnI96-131), 96-139 (TnI96-139), and 96-148 (TnI96-148), to skeletal chicken troponin-C were investigated using multinuclear, multidimensional NMR spectroscopy. The kinetic off-rate and dissociation constants for TnI96-131 (400 s-1, 32 microM), TnI96-139 (65 s-1, <1 microM), and TnI96-148 (45 s-1, <1 microM) binding to TnC were determined from simulation and analysis of the behavior of 1H,15N-heteronuclear single quantum correlation NMR spectra taken during titrations of TnC with these peptides. Two-dimensional 15N-edited TOCSY and NOESY spectroscopy were used to identify 11 C-terminal residues from the 15N-labeled TnI96-148 that were unperturbed by TnC binding. TnI96-139 labeled with 13C at four positions (Leu102, Leu111, Met 121, and Met134) was complexed with TnC and revealed single bound species for Leu102 and Leu111 but multiple bound species for Met121 and Met134. These results indicate that residues 97-136 (and 96 or 137) of TnI are involved in binding to the two domains of troponin-C under calcium saturating conditions, and that the interaction with the regulatory domain is complex. Implications of these results in the context of various models of muscle regulation are discussed.  相似文献   

12.
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   

13.
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   

14.
Yakovlev S  Gorlatov S  Ingham K  Medved L 《Biochemistry》2003,42(25):7709-7716
The beta chain 15-42 sequence of the fibrin(ogen) E region was implicated in heparin binding [Odrljin et al. (1996) Blood 88, 2050-2061]; whether heparin binds to other fibrin(ogen) regions remains to be clarified. To address this question, we studied the interaction of heparin with fibrinogen, fibrin, and their major fragments D(1), D-D, E(1), E(3), and alphaC, which together cover the entire structure of the molecule, by ligand blotting, surface plasmon resonance, and fluorescence. All three techniques revealed that at physiological ionic conditions only fibrin(ogen) and the E(1) fragment bind heparin, indicating that the only physiologically relevant heparin-binding site of fibrin(ogen) is located in its E region. To test whether the beta15-42 sequence is sufficient to form this site or some additional sequences are also involved, we tested the interaction of heparin with a number of beta15-42-containing fragments. The synthetic beta15-42 peptide bound heparin weakly (K(d) = 44.5 microM) while the recombinant beta15-57 and beta15-64 fragments exhibited almost 7-fold higher affinity (K(d) = 6.4 and 7.1 microM, respectively), indicating that the beta43-57 region is also important for heparin binding. At the same time the recombinant dimeric disulfide-linked (beta15-66)(2) fragment which mimics the dimeric arrangement of the beta chains in fibrin bound heparin with high affinity (K(d) = 66 nM), almost 100-fold higher than that for the monomeric fragments. This affinity was similar to those determined for fibrin and the E(1) fragment (K(d) = 72 and 70 nM, respectively) suggesting that (beta15-66)(2) mimics well the heparin-binding properties of the latter two. Altogether, these results indicate that the only heparin-binding site in fibrin(ogen) is formed by NH(2)-terminal portions of the beta chains, including residues beta15-57, and that dimerization is essential for high-affinity binding.  相似文献   

15.
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   

16.
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   

17.
Binding of a synthetic, high-affinity heparin pentasaccharide and of intact heparin to both native and elastase-modified human antithrombin III have been examined by 1H-n.m.r. spectroscopy. The pentasaccharide perturbs many protein resonances in the same way as does intact heparin. There are, however, differences that seem to arise both from fewer contacts in the heparin binding-site when the pentasaccharide binds and from dissimilar conformational changes in the protein. The resonance of the H-2 atom of the histidine, considered to be the N-terminal residue and to be located in the heparin binding-site, is strongly perturbed by heparin binding both to native and modified antithrombin. The pentasaccharide has little effect on this histidine in either protein. Resonances from two of the remaining four histidine units are sensitive to longer-range conformational changes, and show differences between binding of the two heparin species both in native and modified ATIII. It is concluded that the pentasaccharide only partly fills the heparin binding-site and does not produce a conformational change identical to that caused by intact heparin. This is particularly significant as regards the mechanism of action of heparin, because the synthetic pentasaccharide activates ATIII towards Factor Xa, but not towards thrombin.  相似文献   

18.
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   

19.
Previous studies indicated that a major factor in heparin's ability to suppress the proliferation of vascular smooth muscle cells is an interaction with transforming growth factor-beta 1 (TGF-beta 1). Heparin appeared to bind directly to TGF-beta 1 and to prevent the association of TGF-beta 1 with alpha 2-macroglobulin (alpha 2-M). The present studies indicate that 20-70% of iodinated TGF-beta 1 binds to heparin-Sepharose and the retained fraction is eluted with approximately 0.37 M NaCl. Native, unlabelled platelet TGF-beta 1, however, is completely retained by heparin-Sepharose and eluted with 0.9-1.2 M NaCl. Using synthetic peptides, the regions of TGF-beta 1 that might be involved in the binding of heparin and other polyanions were examined. Sequence analysis of TGF-beta 1 indicated three regions with a high concentration of basic residues. Two of these regions had the basic residues arranged in a pattern homologous to reported consensus heparin-binding regions of other proteins. The third constituted a structurally novel pattern of basic residues. Synthetic peptides homologous to these three regions, but not to other regions of TGF-beta 1, were found to bind to heparin-Sepharose and were eluted with 0.15 M-0.30 M NaCl. Only two of these regions were capable of blocking the binding of heparin to 125I-TGF-beta. Immobilization of these peptides, followed by affinity purification of heparin, indicated that one peptide was capable of isolating subspecies of heparin with high and low affinity for authentic TGF-beta 1. The ability of TGF-beta 1 to bind to heparin or related proteoglycans under physiological conditions may be useful in understanding the biology of this pluripotent growth and metabolic signal. Conversely, a subspecies of heparin molecules with high affinity for TGF-beta 1 may be a factor in some of the diverse biological actions of heparin.  相似文献   

20.
Kallistatin is a heparin-binding serine proteinase inhibitor (serpin), which specifically inhibits human tissue kallikrein by forming a covalent complex. The inhibitory activity of kallistatin is blocked upon its binding to heparin. In this study we attempted to locate the heparin-binding site of kallistatin using synthetic peptides derived from its surface regions and by site-directed mutagenesis of basic residues in these surface regions. Two synthetic peptides, containing clusters of positive-charged residues, one derived from the F helix and the other from the region encompassing the H helix and C2 sheet of kallistatin, were used to assess their heparin binding activity. Competition assay analysis showed that the peptide derived from the H helix and C2 sheet displayed higher and specific heparin binding activity. The basic residues in both regions were substituted to generate three kallistatin double mutants K187A/K188A (mutations in the F helix) and K307A/R308A and K312A/K313A (mutations in the region between the H helix and C2 sheet), using a kallistatin P1Arg variant as a scaffold. Analysis of these mutants by heparin-affinity chromatography showed that the heparin binding capacity of the variant K187A/K188A was not altered, whereas the binding capacity of K307A/R308A and K312A/K313A mutants was markedly reduced. Titration analysis with heparin showed that the K312A/K313A mutant has the highest dissociation constant. Like kallistatin, the binding activity of K187A/K188A to tissue kallikrein was blocked by heparin, whereas K307A/R308A and K312A/K313A retained significant binding and inhibitory activities in the presence of heparin. These results indicate that the basic residues, particularly Lys(312)-Lys(313), in the region between the H helix and C2 sheet of kallistatin, comprise a major heparin-binding site responsible for its heparin-suppressed tissue kallikrein binding.  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