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.

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   

2.
The interference of the heparin-neutralizing plasma component S protein (vitronectin) (Mr = 78,000) with heparin-catalyzed inhibition of coagulation factor Xa by antithrombin III was investigated in plasma and in a purified system. In plasma, S protein effectively counteracted the anticoagulant activity of heparin, since factor Xa inhibition was markedly reduced in comparison to heparinized plasma deficient in S protein. Using purified components in the presence of heparin, S protein induced a concentration-dependent reduction of the inhibition rate of factor Xa by antithrombin III. This resulted in a decrease of the apparent pseudo-first order rate constant by more than 10-fold at a physiological ratio of antithrombin III to S protein. S protein not only counteracted the anticoagulant activity of commercial heparin but also of low molecular weight forms of heparin (mean Mr of 4,500). The heparin-neutralizing activity of S protein was found to be mainly expressed in the range 0.2-10 micrograms/ml of high Mr as well as low Mr heparin. S protein and high affinity heparin reacted with apparent 1:1 stoichiometry to form a complex with a dissociation constant KD = 1 X 10(-8) M as determined by a functional assay. As deduced from dot-blot analysis, direct interaction of radiolabeled heparin with S protein revealed a dissociation constant KD = 4 X 10(-8) M. Heparin binding as well as heparin neutralization by S protein increased significantly when reduced/carboxymethylated or guanidine-treated S protein was employed indicating the existence of a partly buried heparin-binding domain in native S protein. Radiolabeled heparin bound to the native protein molecule as well as to a BrCN fragment (Mr = 12,000) containing the heparin-binding domain as demonstrated by direct binding on nitrocellulose replicas of sodium dodecyl sulfate-polyacrylamide gels. Kinetic analysis revealed that the heparin neutralization activity of S protein in the inhibition of factor Xa by antithrombin III could be mimicked by a synthetic tridecapeptide from the amino-terminal portion of the heparin-binding domain. These data provide evidence that the heparin-binding domain of S protein appears to be unique in binding to heparin and thereby neutralizing its anticoagulant activity in the inhibition of coagulation factors by antithrombin III. The induction of heparin binding and neutralization may be considered a possible physiological mechanism initiated by conformational alteration of the S protein molecule.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

3.
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   

4.
Preparations of low molecular weight porcine heparin with an average specific anticoagulant activity of 94 units/mg were fractionated into "active" and "relatively inactive" forms of the mucopolysaccharide of approximately 6000 daltons each. The active fraction was further subdivided into various species with descending but significant affinities for the protease inhibitor as well as decreasing but substantial anticoagulatn potencies. "Highly active" heparin (approximately 8% of the low molecular weight pool) possesses a specific anticoagulant activity of 350 +/- 10 units/mg. The relatively inactive fraction (67% of the low molecular weight pool) exhibits a specific anticoagulant activity of 4 +/- 1 units/mg. The binding of highly active heparin to antithrombin is accurately described by a single-site binding model with a KHep-ATDISS of approximately 1 X 10(-7) M. Variations in this binding parameter secondary to changes in environmental variables indicate that charge-charge interactions as well as an increase in entropy are critical to the formation of the highly active heparin-antithrombin complex. The interaction of relatively inactive heparin with the protease inhibitor is characterized by an apparent KHep-ATDISS of 1 X 10(-4) M. In large measure, this is due to small amounts of residual active mucopolysaccharide (0.5%). The ability of the highly active heparin to accelerate the thrombin-antithrombin interaction was also examined. We were able to demonstrate that the mucopolysaccharide acts as a catalyst in this process and is able to initiate multiple rounds of enzyme-inhibitor complex formation. The rate of enzyme neutralization is increased to a maximum of 2300-fold as the concentration of heparin is raised until the inhibitor is saturated with mucopolysaccharide. Further increases in heparin concentration result in a reduction in the speed of enzyme neutralization. This appears to be due to the formation of thrombin-heparin complexes. A mathematical model is given which provides a relationship between the initial velocity of enzyme neutralization and reactant concentrations.  相似文献   

5.
The specificity of endothelial binding sites for heparin was investigated with heparin fractions and fragments differing in their Mr, charge density and affinity for antithrombin III, as well as with heparinoids and other anionic polyelectrolytes (polystyrene sulphonates). The affinity for endothelial cells was estimated by determining I50 values in competition experiments with 125I-heparin. We found that affinity for endothelial cells increases as a function of Mr and charge density (degree of sulphation). Binding sites are not specific receptors for heparin. Other anionic polyelectrolytes, such as pentosan polysulphates and polystyrene sulphonates, competed with heparin for binding to endothelial cells. Fractions of standard heparin with high affinity for antithrombin III also had greater affinity for endothelium. However, these two properties of heparin (affinity for antithrombin III and affinity for endothelial cells) could be dissociated. Oversulphated heparins and oversulphated low-Mr heparin fragments had lower anticoagulant activity and higher affinity for endothelial cells than did their parent compounds. Synthetic pentasaccharides, bearing the minimal sequence for binding to antithrombin III, did not bind to endothelial cells. Binding to endothelial cells involved partial neutralization of heparin. Bound heparin exhibited only 5% and 7% of antifactor IIa and antifactor Xa specific activity, respectively. In the presence of 200 nM-antithrombin III, and in the absence of free heparin, a limited fraction (approx. 30%) of bound heparin was displaced from endothelial cells during a 1 h incubation period. These data suggested that a fraction of surface-bound heparin could represent a pool of anticoagulant.  相似文献   

6.
Heparin binding to rabbit histidine-rich glycoprotein (HRG) was studied in a purified system, allowing for determination of a heparin dissociation constant of approximately 5.5 X 10(-8) M for the interaction with HRG at pH 7.0. The strong interaction between heparin and HRG was demonstrated to be competitive with the binding of both antithrombin and thrombin to the heparin chain. HRG was further tested as a modulator of the anticoagulant activity of heparin by comparing rates of the heparin-catalyzed reaction between antithrombin and thrombin in the presence and absence of added HRG. The heparin-antithrombin-thrombin reaction was modeled using the formalism of a two-substrate enzyme-catalyzed reaction with heparin as the enzyme and HRG analyzed as an enzyme inhibitor. HRG was shown to compete with both antithrombin and thrombin for binding to heparin by this kinetic analysis. Thus, both the kinetic and heparin-binding data indicate that the mechanism by which HRG modulates heparin anticoagulant activity involves competition for heparin with both the inhibitor and the protease. Inhibition by HRG of the heparin-catalyzed reaction was found to be highly dependent on pH, with a sharp increase in inhibition from about 15% to greater than 90% observed as pH was lowered from 7.4 to 7.0. Since little change in the rate of the heparin-catalyzed inhibition of thrombin by antithrombin occurs in this pH region, the dramatic change in HRG inhibition seen upon pH titration may reflect increasing ionic interaction between heparin and HRG due to the protonation of histidine residues which occurs in this pH region.  相似文献   

7.
Mechanism of charybdotoxin block of a voltage-gated K+ channel.   总被引:5,自引:1,他引:4       下载免费PDF全文
Charybdotoxin block of a Shaker K+ channel was studied in Xenopus oocyte macropatches. Toxin on rate increases linearly with toxin concentration in an ionic strength-dependent fashion and is competitively diminished by tetraethylammonium. On rate is insensitive to transmembrane voltage and to K+ on the opposite side of the membrane. Conversely, toxin off rate is insensitive to toxin concentration, ionic strength, and added tetraethylammonium but is enhanced by membrane depolarization or K+ (or Na+) in the trans solution. Charge neutralization of charybdotoxin Lys27, however, renders off rate voltage insensitive. Our results argue that block of voltage-gated K+ channels results from the binding of one toxin molecule, so that Lys27 enters the pore and interacts with K+ (or Na+) in the ion conduction pathway.  相似文献   

8.
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   

9.
Synthetic peptides based on amino-acid residues 27-38 of human serum amyloid P component represent a novel type of heparin binders as they do not contain clusters of basic amino acids or other known features associated with protein or peptide heparin binding. Here, we characterize the binding using capillary electrophoresis (CE), surface plasmon resonance (SPR), and isothermal titration calorimetry (ITC). By CE, heparin-binding activity was readily apparent for both a regular peptide and a slightly N-terminally modified form, while a sequence-scrambled peptide had no measurable binding. Dissociation constants in the 1-15 microm range were estimated, but only a minor part of the binding isotherm was covered by the experiments. SPR measurements using immobilized peptides verified heparin binding, the range of the binding constants, and the reduced binding of the sequence-scrambled peptide. Structurally defined heparin oligosaccharides were used to establish that while the tetrasaccharide is too small to exhibit strong binding, little difference in binding strength is observed between hexa- and tetradeca-saccharides. These experiments also confirmed the almost complete lack of activity of the sequence-scrambled peptide. The amino-acid sequence-dependent binding and the importance of a disulfide bond in the peptide were verified by ITC, but the experimental conditions had to be modified because of peptide precipitation and ITC yielded significantly weaker binding constants than the other methods. While the precise function of the peptide in the intact protein remains unclear, the results confirm the specificity of the glycosaminoglycan interaction with regard to peptide sequence by applying two additional biophysical techniques and showing that the N-terminal part of the peptide may be modified without changing the heparin binding capabilities.  相似文献   

10.
We have produced a molecule comprising of permanently-activated covalently linked antithrombin and heparin (ATH). This study was designed to elucidate the covalent linkage point(s) for heparin on antithrombin and conformational properties of the ATH molecule. ATH was produced using Schiff base/Amadori rearrangement by incubating antithrombin with unfractionated heparin for 14 d at 40 degrees C. ATH was then digested using Proteinase K, and the heparin-peptide was reacted with NaIO4/NaBH4/mild acid to degrade the heparin moiety. Sequencing of the remaining peptide was performed by Edman degradation with linkage point confirmation by LC-MS. The degree of insertion of the reactive center loop (RCL) of antithrombin into the A-sheet of ATH was examined using synthesized antithrombin RCL peptides. Binding between the peptides and ATH, and the formation of ATH in the presence of the peptides were tested. CD was used to further examine the secondary and tertiary structures of ATH. The results suggest that heparin is conjugated to the amino terminal of antithrombin in the majority of ATH molecules, proximal to the previously determined heparin binding domain of antithrombin. From the linkage data, a model is proposed for the structure of ATH. Studies using the RCL peptides and CD analysis of ATH support this model.  相似文献   

11.
A mouse monoclonal antibody, ST-1, was raised against heparin complexed to Salmonella minnesota. Characterization of this antibody showed that it recognizes an epitope in the intact molecule of heparin that is present regardless of its source or anticoagulant activity. ST-1 is the first monoclonal antibody specific for the intact unmodified molecule of heparin to be described. 3H-labeled heparin in solution was immunoprecipitated by ST-1, and the formation of the 3H-labeled immunocomplex was selectively inhibited by unlabeled heparin. No cross-reactivity of ST-1 was observed with other glycosaminoglycans such as heparan sulfate, chondroitin sulfate, hyaluronic acid, dermatan sulfate, and keratan sulfate, or with polyanionic polymers such as dextran sulfate. Selective removal of the N-sulfate groups or N,O-desulfation of heparin strongly reduced the binding of ST-1. Inhibition of binding was also observed after carbodiimide reduction of the carboxyl groups of the uronic acid units of heparin. Competitive assays of ST-1 binding to heparin immobilized on poly-L-lysine-coated plates using oligosaccharides of different sizes that arose from HNO2 cleavage of heparin showed that the minimum fragment required for reactivity of ST-1 is a decasaccharide.  相似文献   

12.
A synthetic heptadecasaccharide, comprising an antithrombin III binding domain, a thrombin binding domain, and a neutral methylated hexasaccharide sequence, was obtained through a convergent synthesis. This compound displayed in vitro anticoagulant properties similar to that of standard heparin but, in contrast with heparin, escaped neutralization by platelet factor 4, a protein released by activated platelets.  相似文献   

13.
The oxidative mechanism whereby heparin may interact with various proteins was investigated in detail in this work by addressing the role of doses of heparin on the nature and effects of its binding to bovine trypsin, taken as reference protein. Unfractionated heparin was used at concentrations ranging from 6 to 400 microg/ml with a fixed trypsin concentration (250 microg/ml). At concentrations of up to 60 microg/ml, equivalent to trypsin/heparin molar ratios of between 30 and 3, increasing inhibition of amidolytic activity and radical-dependent peptide bond cleavage of the enzyme was observed, with the appearance in the electrophoretic pattern of new bands of trypsin fragments to which heparin was demonstrated to be bound specifically. Structural modifications were also revealed by increases in fluorescence emission spectra. On the whole, however, the alterations induced by these heparin concentrations only involved a limited number of trypsin molecules. At concentrations from 120 to 400 microg/ml (equivalent trypsin/heparin molar ratios of 1.5-0.46), heparin binding to trypsin appeared to cause more profound and generalized alterations of enzyme structure and function, with dose-dependent quenching of fluorescence emission and almost complete loss of amidolytic activity, although evidence of radical production was lacking. Collectively, the results stress the crucial role of heparin dose on both the nature and effects of its binding to trypsin. The change in heparin effects which reflects distinct underlying molecular mechanisms occurs dramatically at a critical concentration threshold. While a specific, radical-generating mechanism operates at low concentrations, less specific ionic linkages, apparently independent of radical production, best explain the effects of high heparin concentrations.  相似文献   

14.
Human von Willebrand factor, a plasma glycoprotein which plays a critical role in regulating hemostasis, binds heparin, but the physiological importance and mode of this interaction is poorly understood. Using the motif of an amino acid sequence of a consensus heparin binding synthetic peptide, a 23-residue sequence (Tyr565-Ala587) of human von Willebrand factor was identified that retains the consensus motif and binds heparin with affinity comparable with native von Willebrand factor and the consensus peptide. In a fluid phase binding assay, the Tyr565-Ala587 peptide competed effectively with von Willebrand factor for binding heparin. Synthesis and testing of peptides overlapping Tyr565-Ala587, as well as adjacent cationic regions, showed this core sequence to be the optimal linear binding domain. Far ultraviolet circular dichroism spectrometry of the Tyr565-Ala587 peptide suggested that the peptide undergoes conformational change upon binding heparin. The Tyr565-Ala587 peptide thus encompasses part (or all) of a functionally important heparin binding domain of von Willebrand factor. Further study of this and related peptides may be useful for exploring how heparin may influence von Willebrand factor-mediated platelet hemostasis.  相似文献   

15.
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   

16.
The effect of salt on the binding of the antimicrobial peptide magainin to POPC lipid bilayers is studied by 40-50 ns molecular dynamics simulations of a POPC bilayer in the presence of different concentrations of Na+ and Cl− ions, corresponding to effective concentrations of 0, 100, 150, 200, 250 and 300 millimolar NaCl, with and without a single molecule of antimicrobial peptide magainin. Simulations without magainin showed that increasing salt concentration leads to the decrease in the area per lipid, a decrease in the head group tilt of the lipids, as well as increased order of lipid tails, in agreement with other recent simulations. Simulations with magainin show that peptide binding to the lipids is stronger at lower concentrations of salt. The peptides disorder the lipids in their immediate vicinity, but this effect is diminished as the salt concentration increases. Our studies indicate that while 50 ns simulations give information on peptide hydrogen bonding and lipid tail ordering that is insensitive to the initial peptide orientation, this run time is not sufficient to equilibrate the peptide position and orientation within the bilayer.  相似文献   

17.
The effect of salt on the binding of the antimicrobial peptide magainin to POPC lipid bilayers is studied by 40-50 ns molecular dynamics simulations of a POPC bilayer in the presence of different concentrations of Na+ and Cl- ions, corresponding to effective concentrations of 0, 100, 150, 200, 250 and 300 millimolar NaCl, with and without a single molecule of antimicrobial peptide magainin. Simulations without magainin showed that increasing salt concentration leads to the decrease in the area per lipid, a decrease in the head group tilt of the lipids, as well as increased order of lipid tails, in agreement with other recent simulations. Simulations with magainin show that peptide binding to the lipids is stronger at lower concentrations of salt. The peptides disorder the lipids in their immediate vicinity, but this effect is diminished as the salt concentration increases. Our studies indicate that while 50 ns simulations give information on peptide hydrogen bonding and lipid tail ordering that is insensitive to the initial peptide orientation, this run time is not sufficient to equilibrate the peptide position and orientation within the bilayer.  相似文献   

18.
N-desulphated heparin, partially N-acylated on average with three oleoyl chains per molecule, inhibits the amidolytic activity of plasmin (IC50 16 nM) and urokinase (IC50 10mM) when assayed on N-p-tosyl-Gly-Pro-Lys-4-nitroanilide and benzoyl-Ala-Gly-Arg-4-nitroanilide substrates respectively. N-desulphated heparin is not inhibitory. This effect requires the covalent binding of oleoyl residues to heparin and it decreases with increasing concentration of Tris-HCl and non-ionic detergents.  相似文献   

19.
A primary heparin-binding site in vitronectin has been localized to a cluster of cationic residues near the C terminus of the protein. More recently, secondary binding sites have been proposed. In order to investigate whether the binding site originally identified on vitronectin functions as an exclusive and independent heparin-binding domain, solution binding methods have been used in combination with NMR and recombinant approaches to evaluate ligand binding to the primary site. Evaluation of the ionic strength dependence of heparin binding to vitronectin according to classical linkage theory indicates that a single ionic bond is prominent. It had been previously shown that chemical modification of vitronectin using an arginine-reactive probe results in a significant reduction in heparin binding (Gibson, A., Baburaj, K., Day, D. E., Verhamme, I. , Shore, J. D., and Peterson, C. B. (1997) J. Biol. Chem. 272, 5112-5121). The label has now been localized to arginine residues within the cyanogen bromide fragment-(341-380) that contains the primary heparin-binding site on vitronectin. One- and two-dimensional NMR on model peptides based on this primary heparin-binding site indicate that an arginine residue participates in the ionic interaction and that other nonionic interactions may be involved in forming a complex with heparin. A recombinant polypeptide corresponding to the C-terminal 129 amino acids of vitronectin exhibits heparin-binding affinity that is comparable to that of full-length vitronectin and is equally effective at neutralizing heparin anticoagulant activity. Results from this broad experimental approach argue that the behavior of the primary site is sufficient to account for the heparin binding activity of vitronectin and support an exposed orientation for the site in the structure of the native protein.  相似文献   

20.
Accumulating evidence indicates that the formation of tumor cell platelet emboli complexes in the blood stream is a very important step during metastases and that the anti-metastasis effects of heparin are partially due to a blockade of P-selectin on platelets. In this study, heparin and chemically modified heparins were tested as inhibitors of three human colon carcinoma cell lines (COLO320, LS174T, and CW-2) binding to P-selectin, adhering to CHO cells expressing a transfected human P-selectin cDNA, and adhering to surface-anchored platelets expressing P-selectin under static and flow conditions. The aim was to screen for heparin derivatives with high anti-adhesion activity but negligible anticoagulant activity. In this study, four modified heparins with high anti-adhesion activity were identified including RO-heparin, CR-heparin, 2/3ODS-heparin, and N/2/3DS-heparin. NMR analysis proved the reliability of structure of the four modified heparins. Our findings suggested that the 6-O-sulfate group of glucosamine units in heparin is critical for the inhibition of P-selectin-mediated tumor cell adhesion. Heparan sulfate-like proteoglycans on these tumor cell surfaces are implicated in adhesion of the tumor cells to P-selectin. Some chemically modified heparins with low anticoagulant activities, such as 2/3ODS-heparin, may have potential value as therapeutic agents that block P-selectin-mediated cell adhesion and prevent tumor metastasis.  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