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.
Specific sequences in heparin are responsible for its modulation of the biological activity of proteins. As part of a program to characterize heparin-peptide and heparin-protein binding, we are studying the interaction of chemically discrete heparin-derived oligosaccharides with peptides and proteins. We report here the isolation and characterization, by one- and two-dimensional 1H NMR spectroscopies, of ten hexasaccharides, one pentasaccharide, and one octasaccharide serine that were isolated from depolymerized porcine intestinal mucosal heparin. Hexasaccharides were chosen for study because they fall within the size range, typically tetra- to decasaccharide in length, of heparin sequences that modulate the activity of proteins. The depolymerization reaction was catalyzed by heparinase I (EC 4.2.2.7) in the presence of histamine, which binds site specifically to heparin. Histamine increases both the rate and extent of heparinase I-catalyzed depolymerization of heparin. It is proposed that oligosaccharides produced by heparinase I-catalyzed depolymerization can inhibit the enzyme by binding to the imidazolium group of histidine-203, which together with cysteine-135 forms the catalytic domain of heparinase I. The increased rate and extent of depolymerization are attributed to competitive binding of the oligosaccharides by histamine.  相似文献   

2.
In order to ascertain which residues in heparin may be responsible for its metal binding capacities we have investigated metal binding to some of its component monosaccharides by 1H and 13C NMR. The diamagnetic Zn ion and the paramagnetic Ni ion were used as probes. 4-Methylumbelliferyl-2-deoxy-2-acetamido-6-O-sulpho-D-glucosamine was used as a model for O-sulphates. Only weak interactions with the sulphate group were found. The 4C1 ring conformation of sodium methyl-beta-D-glucopyranosiduronate was not perturbed by binding to its carboxylate and little evidence exists for chelation. By contrast, the ring conformation of the sodium methyl-alpha-L-idopyranosiduronate is affected by the addition of Zn greater than Pb greater than Cd greater than Ca much greater than K ions. The sodium salt is suggested to be an equilibrium mixture of the 2SO and 1C4 ring conformations. Cation binding to the carboxylate group shifts this equilibrium towards the 1C4 conformation and suggests additional binding to O5 or, less likely, O4. This effect appears to be electrostatic in nature, as excess Na and protonation produce similar shifts. Lead complexation is different from the other ions and suggests some covalent character. The control of the ring conformation of iduronic acid by metal ions may have biological implications for the action of heparin and heparin-like compounds.  相似文献   

3.

Background

Although protamine is effective as an antidote of heparin, there is a need to replace protamine due to its side effects. HIP peptide has been reported to neutralize the anticoagulant activity of heparin. The interaction of HIP analog peptides with heparin and heparin-derived oligosaccharides is investigated in this paper.

Methods

Seven analogues of the heparin-binding domain of heparin/heparan sulfate-interacting protein (HIP) were synthesized, and their interaction with heparin was characterized by heparin affinity chromatography, isothermal titration calorimetry, and NMR.

Results

NMR results indicate the imidazolium groups of the His side chains of histidine-containing Hip analog peptide interact site-specifically with heparin at pH 5.5. Heparin has identical affinities for HIP analog peptides of opposite chirality. Analysis by counterion condensation theory indicates the peptide AC-SRPKAKAKAKAKDQTK-NH2 makes on average ∼ 3 ionic interactions with heparin that result in displacement of ∼ 2 Na+ ions, and ionic interactions account for ∼ 46% of the binding free energy at a Na+ concentration of 0.15 M.

Conclusions

The affinity of heparin for the peptides is strongly dependent on the nature of the cationic side chains and pH. The thermodynamic parameters measured for the interaction of HIP peptide analogs with heparin are strongly dependent on the peptide sequence and pH.

General significance

The information obtained in this research will be of use in the design of new agents for neutralization of the anticoagulant activity of heparin. The site-specific binding of protonated histidine side chains to heparin provides a molecular-level explanation for the pH-dependent binding of β-amyloid peptides by heparin and heparan sulfate proteoglycan and may have implications for amyloid formation.  相似文献   

4.
The interactions between Cu(II) ions and heparin were investigated using several complementary spectroscopic techniques. NMR indicated an initial binding phase involving specific coordination to four points in the structure that recur in slightly different environments throughout the heparin chain; the carboxylic acid group and the ring oxygen of iduronate-2-O-sulfate, the glycosidic oxygen between this residue and the adjacent (towards the reducing end) glucosamine and the 6-O-sulfate group. In contrast, the later binding phase showed little structural specificity. One- and two-dimensional correlated FTIR revealed that complex out of phase (asynchronous) conformational changes also occurred during the titration of Cu(II) ions into heparin, involving the CO and N-H stretches. EPR demonstrated that the environments of the Cu(II) ions in the initial binding phase were tetragonal (with slightly varied geometry), while the later non-specific phases exhibited conventional coordination. Visible spectroscopy confirmed a shift of the absorbance maximum. Titration of Cu(II) ions into a solution of heparin indicated (both by analysis of FTIR and EPR spectra) that the initial binding phase was complete by 15-20 Cu(II) ions per chain; thereafter the ions bound in the non-specific mode. Hetero-correlation spectroscopy (FTIR-CD) improved resolution and assisted assignment of the broad CD features from the FTIR spectra and indicated both in-phase and more complex out of phase (synchronous and asynchronous, respectively) changes in interactions within the heparin molecule during the titration of Cu(II) ions.  相似文献   

5.
The synthetic pentadecapeptide FN-C/H II (KNNQKSEPLIGRKKT-NH(2)) has the sequence of the carboxy-terminal heparin-binding domain of module III(14) of fibronectin. Interaction of FN-C/H II with bovine lung heparin has been studied by (1)H and (23)Na NMR spectroscopy and by heparin affinity chromatography. FN-C/H II binds to heparin from pD <2 up to pD approximately 10; at higher pD, the binding decreases as the lysine side-chain ammonium groups are titrated. Na(+) counterions are displaced from the counterion condensation volume that surrounds sodium heparinate by FN-C/H II, which provides direct evidence that the binding involves electrostatic interactions. The pK(A) values for each of the five ammonium groups of FN-C/H II increase upon binding to heparin which, together with chemical shift data, indicates that the binding involves both delocalized and direct electrostatic interactions between ammonium groups of FN-C/H II and carboxylate and/or sulfate groups of heparin. NMR data also provide evidence for the direct interaction of the guanidinium group of the arginine side chain with anionic sites on heparin. The affinity of heparin for FN-C/H II and for 13 analogue peptides in which lysine and arginine residues were systematically substituted with alanine increases as the number of basic residues increases. The relative contribution of each lysine and arginine to the affinity of heparin for FN-C/H II is R(12) > K(13) > K(14) > K(1) > K(5). Nuclear Overhauser enhancement (NOE) data indicate that, while FN-C/H II is largely unstructured in aqueous solution, the bound peptide interconverts among overlapping, turn-like conformations over the L(9) - T(15) segment of the peptide. NOE data for the interaction of FN-C/H II with a heparin-derived hexasaccharide, together with the number of Na(+) ions displaced from heparin by FN-C/H II as determined by (23)Na NMR, indicates that the peptide binds to a hexasaccharide segment of heparin. Identical NMR and heparin affinity chromatography results were obtained for the interaction of FN-C/H II and its D-amino acid analogue peptide with heparin, which is of interest for the potential use of peptides as therapeutic agents for diseases in which cell adhesion plays a critical role.  相似文献   

6.
Fibroblast growth factors (FGFs) are known to induce formation of new blood vessels, angiogenesis. We show that FGF-induced angiogenesis can be modulated using selectively desulfated heparin. Chinese hamster ovary cells (CHO677) deficient in heparan sulfate biosynthesis were employed to assess the function of heparin/heparan sulfate in FGF receptor-1 (FGFR-1) signal transduction and biological responses. In the presence of FGF-2, FGFR-1 kinase and subsequent mitogen-activated protein kinase Erk2 activities were augmented in a dose-dependent manner, whereas high concentrations of heparin resulted in decreased activity. The length of the heparin oligomer, minimally an 8/10-mer, was critical for the ability to enhance FGFR-1 kinase activity. The N- and 2-O-sulfate groups of heparin were essential for binding to FGF-2, whereas stimulation of FGFR-1 and Erk2 kinases by FGF-2 also required the presence of 6-O-sulfate groups. Sulfation at 2-O- and 6-O-positions was moreover a prerequisite for binding of heparin to a lysine-rich peptide corresponding to amino acids 160-177 in the extracellular domain of FGFR-1. Selectively 6-O-desulfated heparin, which binds to FGF-2 but fails to bind the receptor, decreased FGF-2-induced proliferation of CHO677 cells, presumably by displacing intact heparin. Furthermore, FGF-2-induced angiogenesis in chick embryos was inhibited by 6-O-desulfated heparin. Thus, formation of a ternary complex of FGF-2, heparin, and FGFR-1 appears critical for the activation of FGFR-1 kinase and downstream signal transduction. Preventing complex formation by modified heparin preparations may allow regulation of FGF-2 functions, such as induction of angiogenesis.  相似文献   

7.
Heparan sulfate (HS) chains interact with various growth and differentiation factors and morphogens, and the most interactions occur on the specific regions of the chains with certain monosaccharide sequences and sulfation patterns. Here we generated a library of octasaccharides by semienzymatic methods by using recombinant HS 2-O-sulfotransferase and HS 6-O-sulfotransferase, and we have made a systematic investigation of the specific binding structures for various heparin-binding growth factors. An octasaccharide (Octa-I, DeltaHexA-GlcNSO(3)-(HexA-GlcNSO(3))(3)) was prepared by partial heparitinase digestion from completely desulfated N-resulfated heparin. 2-O- and 6-O-sulfated Octa-I were prepared by enzymatically transferring one to three 2-O-sulfate groups and one to three 6-O-sulfate groups per molecule, respectively, to Octa-I. Another octasaccharide containing 3 units of HexA(2SO(4))-GlcNSO(3)(6SO(4)) was prepared also from heparin. This octasaccharide library was subjected to affinity chromatography for interactions with fibroblast growth factor (FGF)-2, -4, -7, -8, -10, and -18, hepatocyte growth factor, bone morphogenetic protein 6, and vascular endothelial growth factor, respectively. Based upon differences in the affinity to those octasaccharides, the growth factors could be classified roughly into five groups: group 1 needed 2-O-sulfate but not 6-O-sulfate (FGF-2); group 2 needed 6-O-sulfate but not 2-O-sulfate (FGF-10); group 3 had the affinity to both 2-O-sulfate and 6-O-sulfate but preferred 2-O-sulfate (FGF-18, hepatocyte growth factor); group 4 required both 2-O-sulfate and 6-O-sulfate (FGF-4, FGF-7); and group 5 hardly bound to any octasaccharides (FGF-8, bone morphogenetic protein 6, and vascular endothelial growth factor). The approach using the oligosaccharide library may be useful to define specific structures required for binding to various heparin-binding proteins. Octasaccharides with the high affinity to FGF-2 and FGF-10 had the activity to release them, respectively, from their complexes with HS. Thus, the library may provide new reagents to specifically regulate bindings of the growth factors to HS.  相似文献   

8.
We have isolated from nitrous acid cleavage products of heparin two major octasaccharide fragments which bind with high affinity to human antithrombin. Octasaccharide S, with the predominant structure iduronic acid----N-acetylglucosamine 6-O-sulfate----glucuronic acid-----N-sulfated glucosamine 3,6-di-O-sulfate----iduronic acid 2-O-sulfate----N-sulfated glucosamine 6-O-sulfate----iduronic acid 2-O-sulfate----anhydromannitol 6-O-sulfate, is sensitive to cleavage by Flavobacterium heparinase as well as platelet heparitinase and binds to antithrombin with a dissociation constant of (5-15) X 10(-8) M. Octasaccharide R, with the predominant structure iduronic acid 2-O-sulfate----N-sulfated glucosamine 6-O-sulfate----iduronic acid----N-acetylglucosamine 6-O-sulfate----glucuronic acid----N-sulfated glucosamine 3,6-di-O-sulfate----iduronic acid 2-O-sulfate----anhydromannitol 6-O-sulfate, is resistant to degradation by both enzymes and binds antithrombin with a dissociation constant of (4-18) X 10(-7) M. The occurrence of a 15-17% replacement of N-sulfated glucosamine 3,6-di-O-sulfate with N-sulfated glucosamine 3-O-sulfate and a 10-12% replacement of iduronic acid with glucuronic acid in both octasaccharides indicates that these substitutions have little or no effect on the binding of the oligosaccharides to the protease inhibitor. When bound to antithrombin, both octasaccharides produce a 40% enhancement in the intrinsic fluorescence of the protease inhibitor and a rate of human factor Xa inhibition of 5 X 10(5) M-1 s-1 as monitored by stopped-flow fluorometry. This suggests that the conformation of antithrombin in the region of the factor Xa binding site is similar when the protease inhibitor is complexed with either octasaccharide.  相似文献   

9.
K O Badellino  P N Walsh 《Biochemistry》2001,40(25):7569-7580
Inhibition of factor XIa by protease nexin II (K(i) approximately 450 pM) is potentiated by heparin (K(I) approximately 30 pM). The inhibition of the isolated catalytic domain of factor XIa demonstrates a similar potentiation by heparin (K(i) decreasing from 436 +/- 62 to 88 +/- 10 pM) and also binds to heparin on surface plasmon resonance (K(d) 11.2 +/- 3.2 nM vs K(d) 8.63 +/- 1.06 nM for factor XIa). The factor XIa catalytic domain contains a cysteine-constrained alpha-helix-containing loop: (527)CQKRYRGHKITHKMIC(542), identified as a heparin-binding region in other coagulation proteins. Heparin-binding studies of coagulation proteases allowed a grouping of these proteins into three categories: group A (binding within a cysteine-constrained loop or a C-terminal heparin-binding region), factors XIa, IXa, Xa, and thrombin; group B (binding by a different mechanism), factor XIIa and activated protein C; and group C (no binding), factor VIIa and kallikrein. Synthesized peptides representative of the factor XIa catalytic domain loop were used as competitors in factor XIa binding and inhibition studies. A native sequence peptide binds to heparin with a K(d) = 86 +/- 15 nM and competes with factor XIa in binding to heparin, K(i) = 241 +/- 37 nM. A peptide with alanine substitutions at (534)H, (535)K, (538)H, and (539)K binds and competes with factor XIa for heparin-binding in a manner nearly identical to that of the native peptide, whereas a scrambled peptide is approximately 10-fold less effective, and alanine substitutions at residues (529)K, (530)R, and (532)R result in loss of virtually all activity. We conclude that residues (529)K, (530)R, and (532)R comprise a high-affinity heparin-binding site in the factor XIa catalytic domain.  相似文献   

10.
The crystal structure of a dimeric 2:2:2 FGF:FGFR:heparin ternary complex at 3 A resolution has been determined. Within each 1:1 FGF:FGFR complex, heparin makes numerous contacts with both FGF and FGFR, thereby augmenting FGF-FGFR binding. Heparin also interacts with FGFR in the adjoining 1:1 FGF:FGFR complex to promote FGFR dimerization. The 6-O-sulfate group of heparin plays a pivotal role in mediating both interactions. The unexpected stoichiometry of heparin binding in the structure led us to propose a revised model for FGFR dimerization. Biochemical data in support of this model are also presented. This model provides a structural basis for FGFR activation by small molecule heparin analogs and may facilitate the design of heparin mimetics capable of modulating FGF signaling.  相似文献   

11.
To map out the heavy metal binding sites of iduronic acid containing oligosaccharides isolated from human kidneys, we studied Zn(II) binding by nuclear magnetic resonance (NMR) and molecular modeling to two disaccharides isolated after nitrous acid depolymerization of heparin and two synthetic disaccharides representative of the heparin structure, namely, IdopA2S (alpha 1,4)AnManOH, 1 alpha, IdopA2S (alpha 1,4)AnManOH6S, 1b, IdopA2S-(alpha 1,4)GlcNS alpha Me, 2a, and IdopA2S (alpha 1,4)GlcNS6S alpha Me, 2b (see previous article in this series). A conformational analysis of the metal free and metal bound solutions was made by comparing calculated [(NOE)]s, [T1]s, and [J]s to experimental values. The 1C4, 4C1, and 2S0 conformations of the L-idopyranosiduronate ring and the 4E and 4T3 of the anhydro-D-mannitol ring are evaluated as are rotations about the C5-C6 hydroxymethylene of the AnManOH(6S) or GlcNS (6S) residues. The NOE between IdopA2S H1 and H3 and the known NOE between H2 and H5, as well as the T1 of IdopA2S H3, are introduced as NMR observables sensitive to the IdopA2S ring conformation. Similarly, a NOE between IdopA2S H5 and AnManOH(6S) or GlcNS(6S) H3 was observed that directly restricts the allowed interglycosidic conformational space. For all disaccharides, the Zn(II) bound spectral data are consistent with models in which these motions are partially "frozen" such that the 1C4 conformation of the IdopA2S is stabilized along with the 4T3 conformation of the AnManOH(6S) ring. The interglycosidic conformation is also stabilized in one of two minima. Electrostatic potential energy calculations gave the best overall agreement with experiment and suggest metal binding conformations with the carboxylate and ring oxygen of the IdopA2S residues (1C4 conformation) and either O3 of the GlcNS(6S) residues or the sulfate oxygens of the 6-sulphate for 2b providing additional chelating sites. These chelation models concur with the observation of marked 13C and 1H NMR chemical shifts for the IdopA2S resonances and of GlcNS H3 for 2 alpha and GlcNS6S C6 for 2b. This study of model compounds implicates the IdopA2S(alpha 1,4)GlcNS6S group as part of the heavy metal binding site in biologically important acidic oligosaccharides such as heparin.  相似文献   

12.
Although regioselective removal of 6-O-sulfate groups of heparin has been undertaken by several researchers, complete 6-O-desulfation with little side reaction has not been attained successfully. In this work, a modified method with a certain silylating reagent, N-methyl-N-(trimethylsilyl)trifluoroacetamide, has been established to produce completely 6-O-desulfated heparin with few other chemical changes. The degrees of 6-O-desulfation were estimated by means of chemical disaccharide analyses and/or (13)C NMR spectra. Although the completely 6-O-desulfated heparin lost about 20% of 2-O-sulfate groups, any other chemical changes and depolymerization were not detected. The completely 6-O-desulfated heparin displayed strong inhibition of COS-1 cell adhesion to basic fibroblast growth factor (bFGF)-coated well in a dose-dependent manner, as was clarified by the competitive cell-adhesion assay. Furthermore, the completely 6-O-desulfated heparin was shown to promote in vitro A31 fibroblast proliferation in a dose-dependent manner in the presence of bFGF. These results suggest that signal transduction through bFGF/bFGF receptor in A31 cells occurs in the absence of 6-O-sulfate groups in heparin. The involvement of 6-O-sulfate group(s) of heparin/heparan sulfate in the promotion of bFGF mitogenic activity was reported by several groups. This discrepancy between our results and those of other groups would be due to the differences in molecular size of heparin/heparan sulfate derivatives and/or cell species used for the assay.  相似文献   

13.
Mechanism of the anticoagulant action of heparin   总被引:20,自引:0,他引:20  
Summary The anticoagulant effect of heparin, a sulfated glycosaminoglycan produced by mast cells, requires the participation of the plasma protease inhibitor antithrombin, also called heparin cofactor. Antithrombin inhibits coagulation proteases by forming equimolar, stable complexes with the enzymes. The formation of these complexes involves the attack by the enzyme of a specific Arg-Ser bond in the carboxy-terminal region of the inhibitor. The complexes so formed are not dissociated by denaturing solvents, which indicates that a covalent bond may contribute to their stability. This bond may be an acyl bond between the active-site serine of the enzyme and the arginine of the cleaved reactive bond of the inhibitor. However, the native complexes dissociate slowly at near-neutral pH into free enzyme and a modified inhibitor, cleaved at the reactive bond. So, antithrombin apparently functions as a pseudo-substrate that traps the enzyme in a kinetically stable complex.The reactions between antithrombin and coagulation proteases are slow in the absence of heparin. However, optimal amounts of heparin accelerate these reactions up to 2 000-fold, thereby efficiently preventing the formation of fibrin in blood. The accelerating effect, and thus the anticoagulant activity, is shown by only about one-third of the molecules in all heparin preparations, while the remaining molecules are almost inactive. The highly active molecules bind tightly to antithrombin, i.e. with a binding constant of slightly below 108 M–1 at physiological ionic strength, while the relatively inactive molecules bind about a thousand-fold more weakly. The binding of the high-affinity heparin to antithrombin is accompanied by a conformational change in the inhibitor that is detectable by spectroscopic and kinetic methods. This conformational change follows an initial, weak binding of heparin to antithrombin and causes the tight interaction between polysaccharide and inhibitor that is prerequisite to heparin anticoagulant activity. It has also been postulated that the conformational change leads to a more favourable exposure of the reactive site of antithrombin, thereby allowing the rapid interaction with the proteases.Heparin also binds to the coagulation proteases. Recent studies indicate that this binding is weaker and less specific that the binding to antithrombin. Nevertheless, for some enzymes, thrombin, Factor IXa and Factor XIa, an interaction between heparin and the protease, in addition to that between the polysaccharide and antithrombin; apparently is involved in the accelerated inhibition of the enzymes. The effect of this interaction may be to approximate enzyme with inhibitor in an appropriate manner. However, the bulk of the evidence available indicates that binding of heparin to the protease alone cannot be responsible for the accelerating effect of the polysaccharide on the antithrombin-protease reaction.Heparin acts as a catalyst in the antithrombin-protease reaction, i.e. it accelerates the reaction in non-stoichiometric amounts and is not consumed during the reaction. This ability can be explained by heparin being released from the antithrombin-protease complex for renewed binding to antithrombin, once the complex has been formed. Such a decresed affinity of heparin for the antithrombin complex, compared to the affinity for antithrombin alone, has been demonstrated.The structure of the antithrombin-binding region in heparin has been investigated following the isolation of oligosaccharides with high affinity for antithrombin. The smallest such oligosaccharide, an octasaccharide, obtained after partial random depolymerization of heparin with nitrous acid, was found to contain a unique glucosamine-3-O-sulfate group, which could not be detected in other portions of the high affinity heparin molecule and which was absent in heparin with low affinity for antithrombin. The actual antithrombin-binding region within this octasaccharide molecule has been identified as a pentasaccharide sequence with he predominant structure: N-acetyl-D-glucosamine(6-O-SO3)D-glucoronic acidD-glucosamine(N-SO3;3,6-di-O-SO3)L-iduronic acid(2-O-SO3)D-glucosamine(N-SO3;6-O-SO3). In addition to the 3-O-sulfate group, both N-sulfate groups as well as the 6-O-sulfate group of the N-acetylated glucosamine unit appear to be essential for the interaction with antithrombin. The remarkably constant structure of this sequence, as compared to other regions of the heparin molecule, suggests a strictly regulated mechanism of biosynthesis.The ability of heparin to potentiate the inhibition of blood coagulation by antithrombin generally decreases with decreasing molecular weight of the polysaccharide. However, individual coagulation enzymes differ markedly with regard to this molecular-weight dependence. Oligosaccharides in the extreme low-molecular weight range, i.e. octa- to dodecasaccharides, with high affinity for antithrombin have high anti-Factor Xa-activity but are virtually unable to potentiate the inhibition of thrombin. Furthermore, such oligosaccharides are ineffective in preventing experimentally induced venous thrombosis in rabbits. Slightly larger oligosaccharides, containing 16 to 18 monosaccharide residues, show significant anti-thrombin as well as antithrombotic activities, yet have little effect on overall blood coagulation. These findings indicate that the affinity of a heparin fragment for antithrombin is not in itself a measure of the ability to prevent venous thrombo-genesis, and that the anti-Factor Xa activity of heparin is only a partial expression of its therapeutic potential as an antithrombotic agent.The biological role of the interaction between heparin and antithrombin is unclear. In addition to a possible function in the regulation of hemostasis, endogenous heparin may serve as a regulator of extravascular serine proteinases. Mouse peritoneal macrophages have been found to synthesize all the enzymes that constitute the extrinsic pathway of coagulation. Moreover, tissue thromboplastin is produced by these cells in response to a functional interaction with activated T-lymphocytes. The inhibition of this extravascular coagulation system by heparin, released from mast cells, may be potentially important in modulating inflammatory reactions.  相似文献   

14.
Sue SC  Brisson JR  Chang SC  Huang WN  Lee SC  Jarrell HC  Wu W 《Biochemistry》2001,40(35):10436-10446
Glycosaminoglycans (GAGs) have been suggested to be a potential target for cobra cardiotoxin (CTX) with high affinity and specificity via a cationic belt at the concave surface of the polypeptide. The interaction of GAGs, such as high-molecular weight heparin, with CTXs not only can induce aggregation of CTX molecules but also can enhance their penetration into membranes. The binding of short chain heparin, such as a heparin-derived disaccharide [DeltaUA2S(1-->4)-alpha-D-GlcNS6S], to CTX A3 from Taiwan cobra (Naja atra), however, will not induce aggregation and was, therefore, investigated by high-resolution (1)H NMR. A novel heparin binding site on the convex side of the CTX, near the rigid disulfide bond-tightened core region of Cys38, was identified due to the observation of intermolecular NOEs between the protein and carbohydrate. The derived carbohydrate conformation using complete relaxation and conformational exchange matrix analysis (CORCEMA) of NOEs indicated that the glycosidic linkage conformation and the ring conformation of the unsaturated uronic acid in the bound state depended significantly on the charge context of CTX molecules near the binding site. Specifically, comparative binding studies of several heparin disaccharide homologues with two CTX homologues (CTX Tgamma from Naja nigricollis and CTX A3) indicated that the electrostatic interaction of N-sulfate of glucosamine with NH(3)(+)zeta of Lys12 and of the 2-O-sulfate of the unsaturated uronic acid with NH(3)(+)zeta of Lys5 played an important role. These results also suggest a model on how the CTX-heparin interaction may regulate heparin-induced aggregation of the toxin via the second heparin binding site.  相似文献   

15.
The importance of 3-O- and 6-O-sulfated glucosamine residues within the heparin octasaccharide iduronic acid(1)----N-acetylglucosamine 6-O-sulfate(2)----glucuronic acid(3)----N-sulfated glucosamine 3,6-di-O-sulfate(4)----iduronic acid 2-O-sulfate(5)----N-sulfated glucosamine 6-O-sulfate(6)----iduronic acid 2-O-sulfate(7)----anhydromannitol 6-O-sulfate(8) was determined by comparing with synthetic tetra- and penta-saccharides its ability to bind human antithrombin. The octasaccharide had an affinity for antithrombin of 1 X 10(-8) M (10.2 kcal/mol) measured by intrinsic fluorescence enhancement at 6 degrees C. The synthetic pentasaccharide, consisting of residues 2-6, had an affinity of 3 X 10(-8) M (9.6 kcal/mol). The same pentasaccharide, except lacking the 3-O-sulfate on residue 4, had an affinity of 5 X 10(-4) M (4.5 kcal/mol) measured by equilibrium dialysis. The tetrasaccharide, consisting of residues 2-5, bound antithrombin with an affinity of 5 X 10(-6) M (6.8 kcal/mol). The tetrasaccharide, consisting of residues 3-6, had an affinity of 5 X 10(-5) M (5.5 kcal/mol). Since the loss of either the 6-O-sulfated residue 2 or the 3-O-sulfate of residue 4 results in a 4-5 kcal/mol or a 40-50% loss in binding energy of the pentasaccharide, these two residues must be the major contributors to the binding and must be linked to the biologic activity of the octasaccharide.  相似文献   

16.
Vascular endothelial growth factor (VEGF) is a family of glycoproteins with potent angiogenic activity. We reported previously that heparin has an affinity for VEGF165, the major isoform of VEGF, whereas 2-O-desulfated heparin and 6-O-desulfated heparin have weak but significant affinity (Ashikari-Hada, S., Habuchi, H., Kariya, Y., Itoh, N., Reddi, A. H., and Kimata, K. (2004) J. Biol. Chem. 279, 12346-12354). In this study, we first examined the effect of heparin and modified heparins (completely desulfated N-sulfated heparin, 2-O-desulfated heparin, and 6-O-desulfated heparin) on VEGF165-dependent mitogenic activity and tube formation on type I collagen gels of human umbilical vein endothelial cells. Both were enhanced by heparin, but not by modified heparins, suggesting that both the 2-O-sulfate group of hexuronic acid and the 6-O-sulfation group of N-sulfoglucosamine in heparin/heparan sulfate are necessary for VEGF165 activity. We then examined the activation of VEGF receptor (VEGFR) to understand the mechanism. We have made several new findings; 1) heparin yielded a 1.7-fold enhancement of VEGF165-induced phosphorylation of VEGFR-2; 2) depletion of cell surface heparan sulfate by heparinase/heparitinase treatment and preferential reduction of trisulfated disaccharide units of cell surface HS by sodium chlorate treatment resulted in the reduction of such phosphorylation, suggesting the involvement of a heparin-like domain in the phosphorylation of VEGFR-2; and 3) VEGF121, an isoform without the exon 7-encoded region, which has no capacity to bind to heparin, did not show these effects. It is therefore likely that a heparin-like domain of heparan sulfate/heparin forms a complex with VEGF165 and VEGFR-2 via the exon 7-encoded region, thereby enhancing VEGF165-dependent signaling.  相似文献   

17.
We report ELISA studies of the glycosaminoglycan binding properties of recombinant human glial cell line-derived neurotrophic factor (GDNF). We demonstrate relatively high affinity binding as soluble heparin competes with an IC50 of 0.1 micro g/ml. The binding of GDNF to heparin is particularly dependent on the presence of 2-O-sulfate groups. Highly sulfated heparan sulfate is also an effective competitor for GDNF binding. We also show that heparin at low concentrations protects GDNF from proteolytic modification by an endoprotease and also promotes the binding of GDNF to its receptor polypeptide, GFRalpha1. In both of these actions, 2-O-desulfated heparin is less effective. Considered overall, these findings provide strong support for a hypothesis that the bioactivity of GDNF during prenatal development is essentially dependent on the binding of this growth factor to 2-O-sulfate-rich heparin-related glycosaminoglycan.  相似文献   

18.
Disaccharide mimetics of a heparin sequence that binds to fibroblast growth factors were prepared by coupling a D-galactose donor with a methyl beta-D-gluco- or xylopyranoside acceptor. When fully sulfated, the glucose or xylose moieties exist in solution in equilibrium between the (4)C1 and (1)C4 conformers, as confirmed by 1H NMR spectroscopy, thus mimicking the conformationally flexible L-iduronic acid found in heparin. Docking calculations showed that the predicted locations of disaccharide sulfo groups in the binding site of FGF-1 are consistent with the positions observed for co-crystallized heparin-derived oligosaccharides. Predicted binding affinities are in accord with experimental Kd values obtained from binding assays and are similar to the predicted values for a model heparin disaccharide.  相似文献   

19.
BACKGROUND: Membrane-bound ion translocators have important functions in biology, but their mechanisms of action are often poorly understood. Transhydrogenase, found in animal mitochondria and bacteria, links the redox reaction between NAD(H) and NADP(H) to proton translocation across a membrane. Linkage is achieved through changes in protein conformation at the nucleotide binding sites. The redox reaction takes place between two protein components located on the membrane surface: dI, which binds NAD(H), and dIII, which binds NADP(H). A third component, dII, provides a proton channel through the membrane. Intact membrane-located transhydrogenase is probably a dimer (two copies each of dI, dII, and dIII). RESULTS: We have solved the high-resolution crystal structure of a dI:dIII complex of transhydrogenase from Rhodospirillum rubrum-the first from a transhydrogenase of any species. It is a heterotrimer, having two polypeptides of dI and one of dIII. The dI polypeptides fold into a dimer. The loop on dIII, which binds the nicotinamide ring of NADP(H), is inserted into the NAD(H) binding cleft of one of the dI polypeptides. The cleft of the other dI is not occupied by a corresponding dIII component. CONCLUSIONS: The redox step in the transhydrogenase reaction is readily visualized; the NC4 atoms of the nicotinamide rings of the bound nucleotides are brought together to facilitate direct hydride transfer with A-B stereochemistry. The asymmetry of the dI:dIII complex suggests that in the intact enzyme there is an alternation of conformation at the catalytic sites associated with changes in nucleotide binding during proton translocation.  相似文献   

20.
Osteogenin, an extracellular matrix component of bone, is a heparin binding differentiation factor that initiates endochondral bone formation in rats when implanted subcutaneously with an insoluble collagenous matrix. We have examined the interaction of osteogenin with various extracellular matrix components including basement membranes. Osteogenin, purified from bovine bone, binds avidly to type IV collagen and to a lesser extent to both type I and IX collagens. Osteogenin binds equally well to both native and denatured type IV collagen. Both alpha 1 and alpha 2 chains of type IV collagen are recognized by osteogenin. Osteogenin binds to a collagen IV affinity column, and is eluted by 6.0 M urea with 1 M NaCl, pH 7.4, and the eluate contained the osteogenic activity as demonstrated in vivo. Binding of osteogenin to collagen IV is not influenced by either laminin or fibronectin. These results imply that osteogenin binding to extracellular matrix components including collagens I and IV and heparin may have physiological relevance, and such interactions may modulate its local action.  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