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 62 毫秒
1.
Heparin and HS (heparan sulfate) exert their wide range of biological activities by interacting with extracellular protein ligands. Among these important protein ligands are various angiogenic growth factors and cytokines. HS binding to VEGF (vascular endothelial growth factor) regulates multiple aspects of vascular development and function through its specific interaction with HS. Many studies have focused on HS-derived or HS-mimicking structures for the characterization of VEGF165 interaction with HS. Using a heparinase 1-prepared small library of heparin-derived oligosaccharides ranging from hexasaccharide to octadecasaccharide, we systematically investigated the heparin-specific structural features required for VEGF binding. We report the apparent affinities for the association between the heparin-derived oligosaccharides with both VEGF165 and VEGF55, a peptide construct encompassing exclusively the heparin-binding domain of VEGF165. An octasaccharide was the minimum size of oligosaccharide within the library to efficiently bind to both forms of VEGF and a tetradecasaccharide displayed an effective binding affinity to VEGF165 comparable to unfractionated heparin. The range of relative apparent binding affinities among VEGF and the panel of heparin-derived oligosaccharides demonstrate that the VEGF binding affinity likely depends on the specific structural features of these oligosaccharides, including their degree of sulfation, sugar-ring stereochemistry and conformation. Notably, the unique 3-O-sulfo group found within the specific antithrombin binding site of heparin is not required for VEGF165 binding. These findings afford new insight into the inherent kinetics and affinities for VEGF association with heparin and heparin-derived oligosaccharides with key residue-specific modifications and may potentially benefit the future design of oligosaccharide-based anti-angiogenesis drugs.  相似文献   

2.
Heparin was previously reported to potentiate the mitogenic activity of endothelial cell mitogens in a crude extract of bovine hypothalami (Thornton, S. C., Mueller, S. N., and Levine, E. M. (1983) Science 222, 623-625). We and others (Gospodarowicz, D., and Cheng, J. (1986) J. Cell. Physiol. 128, 475-484) have reported that the growth stimulatory effects of acidic fibroblast growth factor (aFGF) are potentiated in a similar manner. We have used these observations as the basis of an assay to characterize the importance of size, sulfation, and anticoagulant activity of heparin in mediating this effect. Partial nitrous acid depolymerization of heparin from porcine intestinal mucosa resulted in a mixture of heparin fragments, containing oligosaccharides ranging from disaccharides to polysaccharides of about 40 monosaccharides in length. This mixture was fractionated by ion exchange chromatography and gel permeation chromatography to obtain size-homogeneous oligosaccharides with different degrees of sulfation. Assay of these heparin-derived saccharides in the presence of a suboptimal concentration of aFGF revealed that a minimum chain length and a certain degree of sulfation is required in order to potentiate the action of aFGF. Low sulfate oligosaccharides (4-16 units) were unable to potentiate aFGF, whereas medium sulfate fractions of octadecasaccharides and larger were able to moderately potentiate aFGF. The potentiation of aFGF by the high sulfate fraction correlated with the saccharide size: 12 or more monosaccharide units were necessary to achieve potentiation equivalent to whole heparin, octa- and decasaccharides were mildly stimulatory, and hexasaccharides were without effect. In the absence of aFGF, intact heparin as well as all the oligosaccharides examined, inhibited the proliferation of capillary endothelial cells to approximately the same degree, between 20 and 50% inhibition. When a tetradecasaccharide was separated into a binding and a nonbinding fraction on matrix-bound antithrombin III, no difference was seen for these fractions in the endothelial cell proliferation assay. These results indicate that both size and sulfation of a heparin-derived oligosaccharide contribute to its ability to interact with aFGF and/or endothelial cells and that this interaction is independent of anticoagulant activity. In addition, our findings suggest that the inhibitory and potentiating effects of heparin on capillary endothelial cells have different structural requirements.  相似文献   

3.
Annexin A2 and heparin bind to one another with high affinity and in a calcium-dependent manner, an interaction that may play a role in mediating fibrinolysis. In this study, three heparin-derived oligosaccharides of different lengths were co-crystallized with annexin A2 to elucidate the structural basis of the interaction. Crystal structures were obtained at high resolution for uncomplexed annexin A2 and three complexes of heparin oligosaccharides bound to annexin A2. The common heparin-binding site is situated at the convex face of domain IV of annexin A2. At this site, annexin A2 binds up to five sugar residues from the nonreducing end of the oligosaccharide. Unlike most heparin-binding consensus patterns, heparin binding at this site does not rely on arrays of basic residues; instead, main-chain and side-chain nitrogen atoms and two calcium ions play important roles in the binding. Especially significant is a novel calcium-binding site that forms upon heparin binding. Two sugar residues of the heparin derivatives provide oxygen ligands for this calcium ion. Comparison of all four structures shows that heparin binding does not elicit a significant conformational change in annexin A2. Finally, surface plasmon resonance measurements were made for binding interactions between annexin A2 and heparin polysaccharide in solution at pH 7.4 or 5.0. The combined data provide a clear basis for the calcium dependence of heparin binding to annexin A2.  相似文献   

4.
The diversity-oriented chemical modification of heparin is shown to afford charge-reduced heparin derivatives that possess increased selectivity for binding heparin-binding proteins. Variable N-desulfonation of heparin was employed to afford heparin fractions possessing varied levels of free amine. These N-desulfonated heparin fractions were selectively N-acylated with structurally diverse carboxylic acids using a parallel synthesis protocol to generate a library of 133 heparin-derived structures. Screening library members to compare affinity for heparin-binding proteins revealed unique heparin-derived structures possessing increased affinity and selectivity for individual heparin-binding proteins. Moreover, N-sulfo groups in heparin previously shown to be required for heparin to bind specific proteins have been replaced with structurally diverse non-anionic moieties to afford identification of charge-reduced heparin derivatives that bind these proteins with equivalent or increased affinity compared to unmodified heparin. The methods described here outline a process that we feel will be applicable to the systematic chemical modification of natural polyanionic polysaccharides and the preparation of synthetic oligosaccharides to identify charge-reduced high affinity ligands for heparin-binding proteins.  相似文献   

5.
The interaction of heparin with heparin binding growth-associated molecule (HB-GAM) was studied using isothermal titration calorimetry (ITC) and surface plasmon resonance (SPR). ITC studies showed that, in solution, heparin bound HB-GAM with a deltaH of -30 kcal/mole corresponding to a dissociation constant (Kd) of 460 nM. The stoichiometry of interaction was 3 moles of HB-GAM per mole of heparin, corresponding to a minimum heparin binding site for HB-GAM of 12-16 saccharide residues. Kinetic measurements of heparin interaction with HB-GAM made by SPR afforded a Kd of 4 nM, suggesting considerably tighter binding when HB-GAM was immobilized on a surface. Affinity chromatography of a sized mixture of heparin oligosaccharides, having a degree of polymerization (dp) of > 14 saccharide units, on HB-GAM-Sepharose demonstrated that oligosaccharides having more than 18 saccharide residues showed the tightest interaction.  相似文献   

6.
Binding and internalization of heparin by vascular smooth muscle cells   总被引:13,自引:0,他引:13  
Previous work from our laboratory has demonstrated that heparin specifically inhibits the proliferation of vascular smooth muscle cells in vivo and in vitro. In this paper, we examine the binding and mode of internalization of heparin by smooth muscle cells. For these studies, radiolabeled and fluoresceinated (FITC) heparin probes were synthesized that retained their antiproliferative capacity. Binding of 3H-heparin to these cells occurs via specific, high-affinity binding sites (Kd = 10(-9) M, 100,000 binding sites per cell). Approximately 80% of the heparin bound to the cell surface was shed into the culture medium within 2 hr. The heparin that was left on the cell surface was internalized with biphasic kinetics. Approximately 50% of the bound material was internalized within 2 hr. After this initial rapid uptake, the rate slowed substantially, with the remaining heparin requiring 1-2 days to be internalized. Binding and uptake of FITC heparin was monitored using video image intensification fluorescence microscopy. When smooth muscle cells were exposed to FITC heparin at 4 degrees C, a diffuse surface staining pattern was observed. After warming the cells to 37 degrees C, intensely fluorescent vesicles were seen superimposed over the diffuse surface staining within 2 min. After 15 min at 37 degrees C, numerous large punctate vesicles were seen inside the cell. After 2 hr these vesicles had concentrated in the perinuclear region. This pattern of uptake, when considered along with the presence of specific, high-affinity binding sites and the initial rapid uptake of 3H-heparin, suggests that heparin enters smooth muscle cells by both receptor-mediated and other endocytic pathways.  相似文献   

7.
Ion-pairing reversed-phased chromatography/mass spectrometry of heparin   总被引:1,自引:0,他引:1  
Heparin and heparin-derived components are widely applied anticoagulant drugs used for amongst other applications medical treatment of deep vein thrombosis and pulmonary embolism. Depolymerisation of native heparin results in complex mixtures of sulfated linear oligosaccharides that are usually not well characterised. In order to further characterise such mixtures, two on-line ion-pairing reverse-phased chromatography electrospray ionisation (ESI) mass spectrometry methods have been developed. One of the systems allows the determination of more than 200 components in a medium molecular weight heparin preparation, whereas the other system can be used to separate isomeric heparin oligosaccharides after previous separation according to size. This latter system allows semi-preparative isolation of isomeric heparin oligosaccharides. The experimental setup includes on-line cation exchange in order to prevent the ion-pairing reagent from entering the mass spectrometer.  相似文献   

8.
The rapid preparation of multimilligram quantities of five heparin-derived oligosaccharides (1–5) is described. These oligosaccharides are the final products obtained from the action of heparin lyase (heparinase, E.C. 4.2.2.7) at its primary sites in the heparin polymer. Five oligosaccharides comprise from 75–85 wt% of commercial porcine mucosal heparins and are recovered in good yield and high purity. Four of these five oligosaccharides were further acted upon at much lower rates by prolonged treatment with heparin lyase or heparan monosulfate lyase (heparitinase, E.C. 4.2.2.8), revealing the subspecificities of these enzymes. These oligosaccharides were used as defined substrates for heparin lyase and heparan monosulfate lyase and their kinetic constants were obtained. Potential applications for these oligosaccharides include their use as defined substrates for purification of heparin monosulfate lyases, and for establishing the catalytic purity of enzyme preparations.  相似文献   

9.
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   

10.
We have characterized the importance of size, sulfation, and anticoagulant activity of heparin in release of basic fibroblast growth factor (bFGF) from the subendothelial extracellular matrix (ECM) and the luminal surface of the vascular endothelium. For this purpose, 125I-bFGF was first incubated with ECM and confluent endothelial cell cultures, or administered as a bolus into the blood of rats, the immobilized 125I-bFGF was then subjected to release by various chemically modified species of heparin and size-homogeneous oligosaccharides derived from depolymerized heparin. Both totally desulfated and N-desulfated heparin failed to release the ECM-bound bFGF. Likewise, substitution of N-sulfate groups of heparin and low molecular weight heparin (fragmin) by acetyl or hexanoyl residues resulted in an almost complete inhibition of bFGF release by these polysaccharides. The presence of O-sulfate groups in heparin increased but was not critical for release of ECM-bound bFGF. Similar structural requirements were identified for release of 125I-bFGF bound to low-affinity sites on the surface of vascular endothelial cells. Oligosaccharides derived from depolymerized heparin and containing as little as 8-10 sugar units were, on a weight basis, equivalent to whole heparin in their ability to release bFGF from ECM. Low-sulfate oligosaccharides were less effective releasers of bFGF as compared to medium- and high-sulfate fractions of the same size oligosaccharides. Heparin fractions with high and low affinity to antithrombin III exhibited a similar high bFGF-releasing activity despite a 200-fold difference in their anticoagulant activities.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Order out of complexity--protein structures that interact with heparin.   总被引:4,自引:0,他引:4  
Many proteins of widely differing functionality and structure are capable of binding heparin. Structural characterisations of the many types of such complexes are being reported in ever-increasing number and at improved resolution. Several crystal structures of complexes formed through the interaction of heparin-derived oligosaccharides with one or more protein partners have been described.  相似文献   

12.
We have previously demonstrated that the Slit proteins, which are involved in axonal guidance and related processes, are high-affinity ligands of the heparan sulfate proteoglycan glypican-1. Glypican-Slit protein interactions have now been characterized in greater detail using two approaches. The ability of heparin oligosaccharides of defined structure (ranging in size from disaccharide to tetradeccasaccharide) to inhibit binding of a glypican-Fc fusion protein to recombinant human Slit-2 was determined using an ELISA. Surface plasmon resonance (SPR) spectroscopy, which measures the interactions in real time, was applied for quantitative modeling of heparin-Slit binding on heparin biochips. Heparin was covalently immobilized on these chips through a pre-formed albumin-heparin conjugate, and the inhibition of Slit binding by heparin, LMW heparin, and heparin-derived oligosaccharides (di-, tetra-, hexa-, and octa-) was examined utilizing solution competition SPR. These competition studies demonstrate that the smallest heparin oligosaccharide competing with heparin binding to Slit was a tetrasaccharide, and that in the ELISA maximum inhibition (approximately 60% at 2 microM concentration) was attained with a dodecasaccharide.  相似文献   

13.
In buffer systems, heparin and low molecular weight heparin (LMWH) directly inhibit the intrinsic factor X-activating complex (intrinsic tenase) but have no effect on the prothrombin-activating complex (prothrombinase). Although chemical modification of LMWH, to lower its affinity for antithrombin (LA-LMWH) has no effect on its ability to inhibit intrinsic tenase, N-desulfation of LMWH reduces its activity 12-fold. To further explore the role of sulfation, hypersulfated LA-LMWH was synthesized (sLA-LMWH). sLA-LMWH is not only a 32-fold more potent inhibitor of intrinsic tenase than LA-LMWH; it also acquires prothrombinase inhibitory activity. A direct correlation between the extent of sulfation of LA-LMWH and its inhibitory activity against intrinsic tenase and prothrombinase is observed. In plasma-based assays of tenase and prothrombinase, sLA-LMWH produces similar prolongation of clotting times in plasma depleted of antithrombin and/or heparin cofactor II as it does in control plasma. In contrast, heparin has no effect in antithrombin-depleted plasma. When the effect of sLA-LMWH on various components of tenase and prothrombinase was examined, its inhibitory activity was found to be cofactor-dependent (factors Va and VIIIa) and phospholipid-independent. These studies reveal that sLA-LMWH acts as a potent antithrombin-independent inhibitor of coagulation by attenuating intrinsic tenase and prothrombinase.  相似文献   

14.
Fluorescence polarization, gel exclusion chromatography and affinity chromatography were used to characterize the interaction of heparins of different size with human plasma fibronectin (Fn) and several of its isolated domains. The fluid-phase interaction of Fn with heparin was dominated by the 30 kDa and 40 kDa Hep-2 domains located near the C-terminal ends of the A and B chains respectively. The 30 kDa Hep-2A domain from the heavy chain was indistinguishable from the 40 kDa Hep-2B domain in this respect; the presence of an additional type III homology unit in the latter had no effect on the binding. Evidence was provided that each Hep-2 domain has two binding sites for heparin. The N-terminal Hep-1 domain reacted weakly in fluid phase even though it binds strongly to immobilized heparin. Fn and Hep-2 fragments were rather undiscriminating in their reaction with fluoresceinamine-labelled heparins of different sizes. However, oligosaccharides smaller than the tetradecasaccharide (14-mer) bound Fn with a 5-10-fold lower affinity. These results suggest that the Hep-2 domains of Fn are able to recognize a broad spectrum of oligosaccharides that presumably vary significantly with respect to the amount and spatial distribution of charge.  相似文献   

15.
We previously reported the identity and purification of two HDL3-binding proteins in rat liver plasma membranes. As these proteins are candidate high density lipoprotein (HDL) receptors and probably multifunctional, including a role in HDL metabolism, we have considerable interest in identifying corresponding proteins that are present in human tissue. This report describes the identification of HDL3-binding sites on human monocytes with the use of fluorescence microscopy and flow cytometry assay. After the incubation of mononuclear cells from human blood with fluorescein isothiocyanate (FITC)-labeled human HDL3, fluorescence micrographs showed dense signals of fluorescent grains on monocytes, but not lymphocytes. A significant increase in FITC intensity on monocytes, but not lymphocytes, was observed by flow cytometry analysis, and the interaction between FITC-HDL3 and human monocytes was concentration-dependent. Although very low density (VLDL) and low density lipoprotein (LDL) were ineffective competitors and HDL2 only partially competed for binding, a 50-fold concentration of HDL3 did compete effectively for binding of FITC-HDL3 to human monocytes. Trypsin treatment reduced the FITC intensity of monocytes, showing that a portion of cell-associated FITC-HDL3 remained bound to the cell surface. Two major HDL-binding proteins were identified in CHAPS-solubilized human mononuclear cells by ligand blotting, using HDL3 as the ligand. Both showed similar binding parameters, specificity, and molecular weight identical to HB1 and HB2 from rat liver plasma membrane. We conclude that corresponding candidate HDL receptors or a similar receptor complex also exist on human blood monocytes.  相似文献   

16.
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   

17.
The therapeutic anticoagulant action of heparin is mediated by the ability of a multifunctional octadecasaccharide region of the molecule to bind to and differentially alter the conformational integrity of antithrombin, and the sugar sequence of the primary binding domain is known. Low ultraviolet circular dichroism spectroscopy of heparin-derived anticoagulant octa-, deca-, dodeca-, tetradeca-, and octadecasaccharides has been useful in elucidating the nature of the sugars that are contained within the second functional domain of the octadecasaccharide. The difference between the spectra of the molar ellipticity of the above sequential oligosaccharides was taken to be the CD spectrum of the corresponding additional disaccharide unit(s). Optical models of the component disaccharides of heparin were derived from CD spectra of heparins having a high degree of sulfation, synthetic glycosides of N-acetylglucosamine, and glycosides of component uronic acids. These were sufficiently distinct in magnitude and spectral position to warrant interpretation of the experimental difference CD spectra. The uronic acids of the disaccharides between deca- and octamer, dodeca- and decamer, and tetradeca- and dodecamer were thereby ascribed to sulfated iduronate, unsulfated iduronate, and glucuronate residues, respectively, while those of the tetrasaccharide between the octadeca- and tetradecasaccharide were tentatively assigned to sulfated iduronate moieties. Interpretation of the difference CD spectra on the basis of the optical models was less certain in regard to the amino sugar components. It appears that the amino sugar derivative between the dodeca- and decamer was N-acetylglucosamine, while the other disaccharides of the octa- to octadecasaccharide probably contained the N-sulfated derivative. A speculative disaccharide sequence drawn from these data indicates that relatively less strongly anionic disaccharides, having nonsulfated uronic acid moieties and N-acetylglucosamine, were flanked by trisulfated disaccharide units, constituting a structural element similar to that which contains the primary binding domain of the anticoagulant.  相似文献   

18.
With the use of [3H]heparin, we recently demonstrated that Leishmania donovani promastigotes express a cell-surface receptor that is specific for the glycosaminoglycan heparin (Mukhopadhyay et al. 1989, The Biochemical Journal, 264, 517-525.). Treatment of the parasite with trypsin abolishes 75-90% of this [3H]heparin-binding activity. When trypsinized promastigotes were resuspended in fresh culture medium in the absence and presence of cycloheximide (10 micrograms/ml), approximately 25-30% of the original heparin-binding capacity was restored within 1 hr, indicating that recruitment of receptors from an internal pool occurred without de novo protein synthesis. Scatchard analysis of the regenerated receptor revealed that the number of regenerated binding sites per cell was 2.3 x 10(5); these sites have a binding affinity of 6.7 x 10(-7) M. Like the native heparin receptors on the surface of freshly isolated cells, the receptors recruited after trypsinization are also highly specific for heparin, as a 25-fold excess of four other glycosaminoglycans displaced less than 10% of bound [3H]heparin from the trypsinized cells. The structural requirements of the ligand heparin, namely the number of monosaccharide units and degree of sulfation, were compared for both the native and regenerated receptor: for both receptors, oversulfated polysaccharide heparin fragments of at least six to eight sugar residues were most efficient at displacing [3H]heparin. The concentrations of oligosaccharide fragments required to displace 50% of [3H]heparin were 0.32 and 0.035 microM for the hexa- and octasaccharides, respectively. Colloidal gold-labeled heparin was bound to promastigotes and visualized by electron microscopy. This analysis revealed that the heparin bound almost exclusively to the flagella of control cells (not subjected to trypsin) and those which had regenerated receptor after trypsinization. The physiological significance of this heparin-binding activity on the surface of promastigotes is discussed.  相似文献   

19.
Tissue factor pathway inhibitor (TFPI) is a heparin-binding protein involved in the extrinsic blood coagulation system. In order to elucidate the minimal size of heparin chain required for the interaction with TFPI, we prepared a series of heparin-derived oligosaccharides with tailored chain length ranged from disaccharide to eicosasaccharide after the successive treatments of heparin, including partial N-desulphation, deaminative cleavage with nitrous acid and gel-filtration. Affinity chromatography study of each oligosaccharide fraction using TFPI as the ligand indicated that increasing the degree of polymerisation causes increased affinity, and that a remarkable change in the affinity occurs between the decamers and dodecamers. Measurement of factor Xa inhibitory activity of TFPI in the presence of each oligosaccharide fraction indicated that the fractions shorter than dodecamers only slightly enhanced the TFPI activity for factor Xa inhibition, while the fractions larger than octadecamers had an effect comparable to full-length heparin. These were compatible to the results from the kinetic analyses of the interaction between TFPI and heparin-derived oligosaccharide with an evanescent wave-based biosensor system, IAsys, using a TFPI C-terminal peptide as the ligand.  相似文献   

20.
C(18) and C(8) bonded stationary phases dynamically coated with cetyltrimethylammonium (CTA) and strong anion exchange (SAX) were developed to obtain separations of oligosaccharide mixtures resulting from chemical or enzymatic depolymerization of heparin. With this method, the retention of sulfated oligosaccharides is directly adjustable depending on the amount of CTA adsorbed into the column. Oligosaccharides containing up to 20 sulfates were separated with a resolving power superior to that of conventional SAX analysis. The stability of the column coating enables hundreds of injections. Using ammonium methane sulfonate aqueous solutions as ultraviolet transparent mobile phases, it was possible to set up double detection, including selective detection of acetylated oligosaccharides. Analytical gel permeation chromatography was directly coupled to CTA-SAX, obtaining a two-dimensional profile of analyzed oligosaccharidic mixtures. A sequencing method of heparin oligosaccharides using partial depolymerization by heparinases according to their size and sulfation pattern and digest analysis by CTA-SAX was developed. A direct application of this method to the analysis of oligosaccharide mixtures obtained by complete digestion of heparins by heparinases I, II, and III was done. It allowed a reliable quantification of heparin building blocks. We also focused our attention on di- and tetrasaccharidic species containing the 3-O-sulfated glucosamines taken as markers of the active sites for antithrombin III. The method was also applied to more complex mixtures resulting from porcine heparin partially depolymerized with heparinase I. The specificity of the reaction was studied up to decasaccharidic fractions.  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