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.
Heparin was modified at carboxyl groups by reaction with several pharmacologically important amino-containing compounds in aqueous medium in the presence of 1-ethyl-3-[3-(dimethylamino)propyl]carbodiimide. In dependence on the nature of the amine and the ratio of reagents, conjugates containing 36–100% amide and 0–25% isoureidocarbonyl groups were synthesized. Isoureidoarylamide groups are present, along with amide moieties, in the products of heparin modification by hydroxyl-containing aromatic amines. The conjugate of heparin with p-aminobenzoic acid contained oligomeric arylamide.  相似文献   

2.
A variety of sulphated polyanions in addition to heparin and dermatan sulphate stimulate the inhibition of thrombin by heparin cofactor II (HCII). Previous investigations indicated that the binding sites on HCII for heparin and dermatan sulphate overlap but are not identical. In this study we determined the concentrations (IC50) of various polyanions required to stimulate thrombin inhibition by native recombinant HCII in comparison with three recombinant HCII variants having decreased affinity for heparin (Lys-173-->Gln), dermatan sulphate (Arg-189-->His), or both heparin and dermatan sulphate (Lys-185-->Asn). Pentosan polysulphate, sulphated bis-lactobionic acid amide, and sulphated bis-maltobionic acid amide resembled dermatan sulphate, since their IC50 values were increased to a much greater degree (>/=8-fold) by the mutations Arg-189-->His and Lys-185-->Asn than by Lys-173-->Gln (Gln and Lys-185-->Asn (>/=6-fold) than by Arg-189-->His (相似文献   

3.
The modification of hyaluronidase by aldehydodextran regulates inhibition of the enzyme by heparin. A 70-90% modification of the surface amino groups of hyaluronidase results in sharp conformational changes and a substantial decrease of its inhibition by heparin, whereas hyaluronidase derivatives with a modification degree of 96-100% are practically uninhibited.  相似文献   

4.
A study was designed to determine the feasibility of developing in vitro maturation, fertilization and culture systems utilizing follicular oocytes and epididymal spermatozoa collected from llamas at slaughter. From a total of 1324 cumulus oocyte complexes (COCs) recovered, 972 were cultured in 50-ul drops of TCM-199 medium with 10% heat inactivated steer serum (DBS) and hormones for 30 h. After maturation, the oocytes were randomly allocated into 4 groups in a 2x2 factorial design: cumulus-enclosed oocytes, 2 ug/ml heparin (Group 1); cumulus-enclosed oocytes, 5 ug/ml heparin (Group 2); denuded oocytes, 2 ug/ml heparin (Group 3); and denuded oocytes, 5 ug/ml heparin (Group 4). Denuded oocytes were obtained for groups 3 and 4 by vortexing. Epididymides were also collected at slaugther and fresh spermatozoa (for each replicate) were obtained by mincing the cauda epididymis with a scalpel blade. A total of 721 oocytes were inseminated with 2-3 x 10(6) epididymal spermatozoa/ml in a 50-ul drop of FERT-TALP medium. After 18 h of in vitro insemination, 234 oocytes were placed in a llama oviductal epithelial cell (LLOEC) co-culture in TCM-199 for 9 d. All cultures were done at 38.5 degrees C under 5% CO(2) in air with high humidity. The rate of fertilization, initial cleavage and development in co-culture were evaluated and compared. Of 192 oocytes examined for signs of fertilization, 56 (29.2%) were penetrated by spermatozoa with 57.1% (32 56 ) of the penetrated oocytes having a male and female pronucleus. There were no differences among treatment groups in total fertilization. However, the frequency of oocytes fertilized normally tended to be higher in the denuded oocytes 67.7% (21 31 ) than the oocytes inseminated with cumulus cells 44.0% (11 25 ) independent of heparin concentration (P<0.06). The total embryo development rate to the 2 cells to blastocyst stage was 32.1% (75 234 ). There was no difference in development rate between groups. From the 234 oocytes co-cultured in LLOEC for 9 d, 15.8% developed into 2 to 16 cells, 5.6% into morulae, 6.0% into early/expanded blastocysts and 4.7% into hatching/hatched blastocysts. The results indicate that an in vitro fertilization system is possible in the llama utilizing slaughterhouse material and that llama oocytes can be fertilized in the presence of heparin and epididymal spermatozoa.  相似文献   

5.
Binding and endocytosis of heparin by human endothelial cells in culture   总被引:8,自引:0,他引:8  
Binding of heparin and low molecular weight heparin fragments (CY 222, Mr range 1500-8000) to human vascular endothelial cells was studied. Primary culture of human umbilical vein endothelial cells and either 125I or 3H-labeled heparin or [125I]CY 222 were used. Slow, saturable and specific binding was found. No other tested glycosaminoglycan, excepting a highly sulfated heparan fraction, was able to compete for heparin binding. Two groups of binding sites for [3H]heparin could be distinguished: one with high affinity (Kd = 0.12 microM) and another with lower affinity (Kd = 1.37 microM) and a relative large capacity of binding (1.16 X 10(7) molecules/cell) was calculated. The Kd for unlabeled heparin, as calculated from competition experiments, was 0.23 microM. Much lower affinity was calculated for unlabeled low molecular weight heparin fragments CY 222 (Kd = 4.3 microM) from competition experiments with [125I]CY 222. The binding reversibility was only partial for unfractionated heparin. Even by chasing with unlabeled compound, a fraction of 25-30% was not dissociable from endothelial cells. This fraction was much lower if incubation was carried out at 4 degrees C. The addition of basic proteins (histones) to the incubation medium greatly enhanced the undissociable binding at 37 degrees C, but not at 4 degrees C. The undissociable fraction of heparin was not available to degradation by purified microbial heparinase. These results suggest that a fraction of bound heparin is internalized by the vascular endothelium.  相似文献   

6.
The anisotropy of the fluorescence of dansyl (5-dimethylaminonaphthalene-1- sulphonyl) groups covalently attached to human platelet factor 4 was used to detect the macromolecular compounds formed when the factor was mixed with heparin. At low heparin/protein ratios a very-high-molecular-weight compound (1) was formed that dissociated to give a smaller compound (2) when excess heparin was added. 2. A large complex was also detected as a precipitate that formed at high protein concentrations in chloride buffer. It contained 15.7% (w/w) polysaccharide, equivalent to four or five heparin tetrasaccharide units per protein tetramer. In this complex, more than one molecule of protein binds to each heparin molecule of molecular weight greater than about 6 X 10(3).3. The stability of these complexes varied with pH, salt concentration and the chain length of the heparin. The limit complexes found in excess of the larger heparins consisted of only one heparin molecule per protein tetramer, and the failure to observe complexes with four heparin molecules/protein tetramer is discussed.  相似文献   

7.
The antithrombin-binding region of heparin is a pentasaccharide sequence with the predominant structure -GlcNAc(6-OSO3)-GlcA-GlcNSO3(3,6-di-OSO3)-Ido A(2-OSO3)- GlcNSO3(6-OSO3)-. By using the 3-O-sulfated glucosamine residue as a marker for the anti-thrombin-binding sequence, the location of this sequence within the heparin chain was investigated. Heparin with high affinity for antithrombin (HA-heparin) contains few N-acetyl groups located outside the antithrombin-binding region, and cleavage at such groups was therefore expected to be essentially restricted to this region. HA-heparin was cleaved at N-acetylated glucosamine units by partial deacetylation followed by treatment with nitrous acid at pH 3.9, and the resulting fragments with low affinity for anti-thrombin (LA-fragments) were recovered after affinity chromatography on immobilized antithrombin. The LA-fragments were further divided into subfractions of different molecular size by gel chromatography and were then analyzed with regard to the occurrence of the nonreducing terminal GlcA-GlcNSO3(3,6-di-OS-O3)- sequence. Such units were present in small, intermediate-sized as well as large fragments, suggesting that the antithrombin-binding regions were randomly distributed along the heparin chains. In another set of experiments, HA-heparin was subjected to limited, random depolymerization by nitrous acid (pH 1.5), and the resulting reducing terminal anhydromannose residues were labeled by treatment with NaB3H4. The molecular weight distributions of such labeled LA-fragments, determined by gel chromatography, again conformed to a random distribution of the antithrombin-binding sequence within the heparin chains. These results are in apparent disagreement with previous reports (Radoff, S., and Danishefsky, I. (1984) J. Biol. Chem. 259, 166-172; Rosenfeld, L., and Danishefsky, I. (1988) J. Biol. Chem. 263, 262-266) which suggest that the antithrombin-binding region is preferentially located at the nonreducing terminus of the heparin molecule.  相似文献   

8.
Structural features of heparin potentially important for heparanase-inhibitoryactivity were examined by measuring the ability of heparin derivativesto affect the degradation of [3H]acetylated heparan sulphateby tumor cell heparanases. IC50 values were determined usingan assay which distinguished degraded from undegraded substrateby precipitation of the latter with cetylpyridinium chloride(CPC). Removal of heparin's 2-O-sulphate and 3-O-sul-phate groupsenhanced heparanase-inhibitory activity (50%). Removal of itscarboxyl groups slightly lowered the activity (18%), while combiningthe treatments abolished the activity. At least one negativecharge on the iduronic acid/idose moiety, therefore, is necessaryfor heparanase-inhibitory activity. Replacing heparin's N-sulphategroups with N-acetyl groups reduced its activity (37%). Comparingthis heparin derivative with 2,3-O-de-sulphated heparin, theplacement of sulphate groups appears important for activitysince the two structures have similar nominal linear chargedensity. In addition, unsubstituted uronic acids are nonessentialfor inhibition since their modification (periodate-oxidation/borohydride-reduction)enhanced rather than reduced heparanase-inhibitory activity.The most effective heparanase inhibitors (2,3-O-desulphatedheparin, and [periodate-oxidized, borohydride-reduced] heparin)were tested in the chick chorioallantoic membrane (CAM) bioassayfor anti-angiogenic activity and found to be at least as efficaciousas heparin. 2,3-O-desulphated heparin also significantly decreasedthe tumor growth of a subcutaneous human pancreatic (Ca-Pan-2)adenocarcinoma in nude mice and prolonged the survival timesof C57BL/6N mice in a B16-F10 melanoma experimental lung metastasisassay. angiogenesis chemically-modified heparins endoglycosidase hepara sulphate cancer  相似文献   

9.
Cesaretti M  Luppi E  Maccari F  Volpi N 《Glycobiology》2004,14(12):1275-1284
Heparin with high anticoagulant activity (activated partial thromboplastin time of 347 +/- 56.4 and anti-Xa activity of 317 +/- 48.3) was isolated from the marine clam species Tapes phylippinarum in an amount of approximately 2.1 mg/g dry animals. Agarose-gel electrophoresis showed a high content of the slow-moving heparin component (22 +/- 6.8%) and 78 +/- 5.4% of the fast-moving species. An average molecular mass of 13,600 was calculated by PAGE analysis, whereas a number average molecular weight Mn value of 10,700, a weight average molecular weight Mw of 14,900, and a dispersity index Mn/Mw of 1.386 were obtained by high-performance size-exclusion chromatography. Structural analysis of clam heparin, performed by depolymerizing heparin samples with heparinase (EC 4.2.2.7) and then separating the resulting unsaturated oligosaccharides by strong anion exchange-HPLC revealed the presence of large amounts (more than 130% than standard pharmaceutical heparin obtained from bovine intestine) of the oligosaccharide sequence bearing part of the ATIII-binding region, DeltaUA2S (1-->4)-alpha-D-GlcN2S6S (1-->4)-alpha-L-IdoA (1-->4)-alpha-D-GlcNAc6S (1-->4)-beta-D-GlcA (1-->4)-alpha-D-GlcN2S3S6S in the T. phylippinarum heparin, in comparison with bovine mucosal heparin and a sample of porcine mucosal heparin previously published. Furthermore, as expected from the oligosaccharide compositional analysis, due to the presence of a great mol % (80.6%) of the trisulfated disaccharide DeltaUA2S(1-->4)-alpha-D-GlcN2S6S, mollusc heparin is a more sulfated polysaccharide than bovine mucosal heparin (73.5%) and a sample of porcine mucosal (72.8%) heparin previously reported. To our knowledge, this is the first article describing a clam heparin having the ATIII binding site mainly identical to that of human and porcine intestinal mucosal heparins and bovine intestinal mucosal heparin but different from that found in beef lung heparin.  相似文献   

10.
Poly(3-hydroxyoctanoate) (PHO) films were treated with plasma of different discharge powers (10-50 W) and then treated with acryl amide solutions in order to prepare films with surfaces that contained different amounts of amide groups. The surfaces were characterized by contact angle measurement, attenuated total reflectance infrared spectroscopy, electron spectroscopy for chemical analysis, and scanning electron microscopy. Results from all these measurements indicated that amide groups were present on the surfaces. The amount of amide groups increased in proportion to the discharge power of the plasma. The interaction of Chinese hamster ovary cells with these grafted surfaces was investigated. The number of cells that adhered to and grew on the surfaces was highest for films grafted at 30 W of plasma discharge power, indicating that the moderate hydrophilicity was optimal for cells to adhere and grow. The present results support the suggestion that acryl amide-grafted PHO could be used as cell-compatible biomedical applications.  相似文献   

11.
Murayama K  Ozaki Y 《Biopolymers》2002,67(6):394-405
The molten globule-like states of ovalbumin (OVA) in acid aqueous solutions are investigated by generalized two-dimensional (2D) Fourier transform near-IR (FT-NIR) correlation spectroscopy. This new method allows us to explore the changes in hydration and the secondary structure simultaneously. FT-NIR spectra are measured for OVA aqueous solutions with concentrations of 1, 2, 3, 4, and 5 wt % over a pH range of 2.4-5.4. Concentration-perturbed 2D correlation spectra are calculated for the spectra in the 4850-4200 and 7500-5350 cm(-1) regions at different pH values. The 2D NIR synchronous spectrum in the 4850-4200 cm(-1) region shows a significant change upon going from pH 5.4 to 3.6. An autopeak at 4265 cm(-1) that is due to a combination of a symmetric CH(2) stretching mode and a CH(2) bending mode of side chains seen at pH 5.0 disappears completely in the synchronous spectrum at pH 3.6. This suggests that some amino acid residues of OVA are subjected to microenvironmental changes with decreasing pH. More remarkable changes are observed in the synchronous spectra at pHs below 2.8. A band near 4600 cm(-1) arising from a combination of amide B and amide II modes (amide B/II) shifts downward with considerable broadening between pH 3.0 and 2.4, suggesting that the strength of the hydrogen bonds of amide groups of OVA changes significantly. The synchronous and asynchronous spectra in the 4850-4200 cm(-1) region show that the intensities of the bands attributable to amide groups and side chains of OVA and that of the band near 4800 cm(-1) arising from water change in phase with the increase in the concentration above pH 2.8, but they vary out of phase below pH 2.8. The 2D synchronous map in the 7500-5350 cm(-1) region also shows marked changes upon going from pH 2.8 to 2.6. A broad autopeak at around 6950 cm(-1) assigned to free water and bound water with weak hydrogen bonds becomes very weak in the synchronous spectrum at pH 2.6, while broad autopeaks around 6450 cm(-1) suddenly appear that are due to bound water with several hydrogen bonds and the first overtone of an NH stretching mode of the amide groups of OVA. Therefore, it is very likely that protein hydration and the hydrogen bonds of amide groups change simultaneously in a narrow pH region of 2.8-2.6. It is probably that below pH 2.6 the protein assumes a molten globule-like state in which the whole molecule is very flexible, and side chains (but not the backbone chain) fluctuate significantly.  相似文献   

12.
Heparin with high anticoagulant activity was isolated from the two marine clam species Anomalocardia brasiliana and Tivela mactroides. A large portion of the polysaccharide chains of both preparations bound with high affinity to immobilized antithrombin. Titrations monitored by tryptophan fluorescence showed that clam polysaccharide chains with Mr approximately 22,500 contained up to three binding sites for antithrombin and that the binding constants for the interaction of these chains with antithrombin were higher than those reported for mammalian heparin of comparable size. Structural analysis of clam heparin fractions and subfractions of clam heparin with differing affinity for immobilized antithrombin revealed the presence of large amounts (up to 25-30% of the total disaccharide units) of the 3-O-sulfated saccharide sequences (-GlcNSO3)-GlcA-GlcNSO3(3-OSO3)- and (-GlcNSO3)-GlcA-GlcNSO3(3,6-di-OSO3)-, previously identified as unique markers for the antithrombin-binding region of heparin. The content of these saccharide sequences was found to increase with increasing affinity of the parent polysaccharide for antithrombin. Structural analysis of the clam heparins also demonstrated the occurrence of a novel saccharide sequence, tentatively identified as (-GlcNSO3)-IdA-GlcNSO3(3,6-di-OSO3)-, that has not previously been found in heparin or related polysaccharides. The contents of this latter sequence, at most 3-4% of the total disaccharide units, showed no correlation with the affinity for antithrombin.  相似文献   

13.
Heparin binding site affinity and density on human spermatozoa were compared between fertile and infertile men with normal or abnormal results in the zona-free hamster ova-sperm penetration assay (SPA). A portion of fresh semen from fertile donors and potentially infertile men was processed through the SPA while the remainder of the ejaculate was used to quantitate heparin binding on spermatozoa. Saturation binding assays with [3H]heparin (15-375 nM) were analysed for 3 groups of men: (1) infertile patients with abnormal SPA results, (2) infertile patients with normal SPA results and (3) fertile donors. The heparin binding site density was significantly higher in men who possessed normal SPA results (infertile men and fertile donors) than in infertile men with abnormal scores in the SPA. There was no difference in heparin binding affinity between the three groups. These findings suggest that the heparin binding-site density may be related to the ability of human spermatozoa to undergo successfully the acrosome reaction.  相似文献   

14.
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   

15.
We have analysed the hydration of main-chain carbonyl and amide groups in 24 high-resolution well-refined protein structures as a function of the secondary structure in which these polar groups occur. We find that main-chain atoms in beta-sheets are as hydrated as those in alpha-helices, with most interactions involving "free" amide and carbonyl groups that do not participate in secondary structure hydrogen bonds. The distributions of water molecules around these non-bonded carbonyl groups reflect specific steric interactions due to the local secondary structure. Approximately 20% and 4%, respectively of bonded carbonyl and amide groups interact with solvent. These include interactions with carbonyl groups on the exposed faces of alpha-helices that have been correlated previously with bending of the helix. Water molecules interacting with alpha-helices occur mainly at the amino and carbonyl termini of the helices, in which case the solvent sites maintain the hydrogen bonding by bridging between residues i and i-3 or i-4 at the amino terminus and between i and i+3 or i+4 at the carbonyl terminus. We also see a number of solvent-mediated Ncap and Ccap interactions. The water molecules interacting with beta-sheets occur mainly at the edges, in which case they extend the sheet structure, or at the ends of strands, in which case they extend the beta-ladder. In summary, the solvent networks appear to extend the hydrogen-bonding structure of the secondary structures. In beta-turns, which usually occur at the surface of a protein, exposed amide and carbonyl groups are often hydrated, especially close to glycine residues. Occasionally water molecules form a bridge between residues i and i+3 in the turn and this may provide extra stabilization.  相似文献   

16.
In order to study base pairing properties of the amide group in DNA duplexes, a nucleoside analog, 1-(2'-deoxy-beta-D-ribofuranosyl)pyrrole-3-carboxamide, was synthesized by a new route from the ester, methyl 1-(2'-deoxy-3',5'-di-O-p -toluoyl-beta-D-erythro-pentofuranosyl)pyrrole-3-carboxylate, obtained from the coupling reaction between 1-chloro-2-deoxy-3,5-di-O -toluoyl-d-erythropentofuranose and methyl pyrrole-3-carboxylate by treatment with dimethylaluminum amide. 1-(2'-Deoxy-beta-D-ribofuranosyl)pyrrole-3-carboxamide was incorporated into a series of oligodeoxyribonucleotides by solid-phase phosphoramidite technology. The corresponding oligodeoxyribonucleotides with 3-nitropyrrole in the same position in the sequence were synthesized for UV comparison of helix-coil transitions. The thermal melting studies indicate that pyrrole-3-carboxamide, which could conceptually adopt either a dA-like or a dI-like hydrogen bond conformation, pairs with significantly higher affinity to T than to dC. Pyrrole-3-carboxamide further resembles dA in the relative order of its base pairing preferences (T >dG >dA >dC). Theoretical calculations on the model compound N-methylpyrrole-3-carboxamide using density functional theory show little difference in the preference for a syntau versus anti conformation about the bond from pyrrole C3 to the amide carbonyl. The amide groups in both the minimized antitau and syntau conformations are twisted out of the plane of the pyrrole ring by 6-14 degrees. This twist may be one source of destabilization when the amide group is placed in the helix. Another contribution to the difference in stability between the base pairs of pyrrole-3-carboxamide with T and pyrrole-3-carboxamide with C may be the presence of a hydrogen bond in the former involving an acidic proton (N3-H of T).  相似文献   

17.
Periodate-oxidized/borohydride-reduced 2-O-desulfated heparin (OR2DSH) was prepared using intact heparin from pig intestine as the starting material. Successive treatments of the heparin by oxidation with sodium periodate and reduction with sodium borohydride yielded periodate-oxidized/borohydride-reduced heparin (OR-heparin). Subsequent 2-O-desulfation of OR-heparin, according to a previously established method, yielded OR2DSH. Digestion of OR2DSH with heparitinases generated unsaturated disaccharides, comprising 86.5% DeltaDiHS-(6,N)S (DeltaUA1-->4GlcNS(6S)) and 13.5% DeltaDiHS-NS (DeltaUA1-->4GlcNS), as well as undigested oligosaccharides in which uronate moieties were derivatized by the cleavage of the covalent bond between the C-2 and C-3 positions by periodate-oxidation. The molecular mass of OR2DSH was determined to be 11 kDa, which is almost the same as those of other heparin derivatives such as 2-O-desulfated heparin (2DSH), 6-O-desulfated heparin (6DSH) and N-desulfated N-reacetylated heparin (NDSNAc-heparin). The ability of OR2DSH to enhance neurite outgrowth-promoting activity was evaluated using the explant culture of neocortical tissue from rat embryo in which endogenous heparan sulfate at the cell surface lost substantial numbers of sulfate groups by the action of 40 micro M sodium chlorate. The maximum activity of OR2DSH (29.7%) was achieved at 10 micro g/ml, and those of OR-heparin (21.7%), 2DSH (18.7%) and intact heparin (16.3%) were 100 micro g/ml, whereas that of NDSNAc-heparin (16.5%) was 1,000 micro g/ml. Completely 6-O-desulfated heparin (100:6DSH) exhibited very weak activity (3.3%) at 1,000 micro g/ml. These results suggest that the potency of OR2DSH to enhance neurite outgrowth-promoting activity is exerted synergetically by two different components in OR2DSH, i.e., the IdoA alpha1-->4GlcNS(6S) unit, which contains 6-O- and 2-N-sulfate groups, and the uronate moiety in which the covalent bond between C-2 and C-3 is cleaved, although the mode of action remains to be clarified.  相似文献   

18.
In investigating the role of cell-extracellular matrix interactions in cell adhesion and growth control, the effects of heparin on cell-collagen interactions were examined. Exponentially growing Balb/c-3T3 fibroblasts were radiolabelled with 3H-thymidine and detached from tissue culture surfaces using EDTA, and cell attachment to various types of collagen substrata was assayed in the presence or absence of heparin or other glycosaminoglycans (GAGs) or dextran sulfate (40 K). Cells attached readily (70-90%) to films of types I and V, but not to type III collagen. The number of cells bound to types I and V collagen films was inhibited by 10-50% when heparin was present from 0.1-100 micrograms/ml. Cell-collagen attachment was also inhibited by dextran sulfate, and to a lesser extent by dermatan sulfate, but chondroitin sulfates A and C and hyaluronic acid showed no effect. Heparin was active even at early time points in the adhesion assay, suggesting it may disrupt cell-collagen attachment. To study the effects of heparin in modulating cell growth on collagen, growth arrested cells cultured on type I collagen films were serum stimulated in the presence of heparin or other GAGs for 3 days. Growth was inhibited (greater than 40%) only by heparin and dextran sulfate. Interaction of heparin fragments (Mr less than or equal to 6KD) with type I collagen was analyzed by affinity co-electrophoresis (Lee and Lander, 1991) and showed higher affinity heparin binding to native as compared with denatured collagen. These data suggest that sites within native collagen may mediate Balb cell-collagen and heparin-collagen interactions, and such interactions may be relevant towards understanding heparin's antiproliferative activity in vivo and in vitro.  相似文献   

19.
The effects of heparin on ion channels formed by Staphylococcus aureus alpha-toxin (ST channel) in lipid bilayers were studied under voltage clamp conditions. Heparin concentrations as small as 100 pM induced a sharp dose-dependent increase in channel voltage sensitivity. This was only observed when heparin was added to the negative-potential side of lipid bilayers in the presence of divalent cations. Divalent cations differ in their efficiency: Zn2+>Ca2+>Mg2+. The apparent positive gating charge increased 2-3-fold with heparin addition as well as with acidification of the bathing solution. 'Free' carboxyl groups and carboxyl groups in ion pairs of the protein moiety are hypothesized to interact with sulfated groups of heparin through divalent cation bridges. The cis mouth of the channel (that protrudes beyond the membrane plane on the side of ST addition and to which voltage was applied) is less sensitive to heparin than the trans-mouth. It is suggested that charged residues which interact with heparin at the cis mouth of ST channels and which contribute to the effective gating charge at negative voltage may be physically different from those at the trans mouth and at positive voltage.  相似文献   

20.
Nuclear magnetic resonance spectroscopic studies of the strain-specific secondary cell wall polymer (SCWP) of the Gram-positive, moderately thermophilic organism Geobacillus tepidamans GS5-97T reveal two glycoforms consisting of identical tetrasaccharide repeating units with different chemical modifications of the amide moieties. On the basis of sugar analyses along with 1D and 2D 1H, 13C, 15N, and 31P NMR spectroscopy at natural isotope abundance, the basic backbone structure of the SCWP was established to be [beta-D-Manp-2,3-diNAcANH2-(1-->6)-alpha-D-Glcp-(1-->4)-beta-D-Manp-2,3-diNAcANH2-(1-->3)-alpha-D-GlcpNAc-(1-->]6-(1-->O)-PO2-(O-->6)-MurNAc-, with modifications of the amide groups. In one glycoform, all beta-D-Manp-2,3-diNAcANH2 (2,3-diacetamido-2,3-dideoxy-beta-D-mannopyranuronamide, ManpANH2) residues are substituted with two acetyl groups (glycoform I) at the amide group at C-6; in the other glycoform (glycoform II), only one proton of this amide group is substituted by an acetyl group. The ratio between both the glycoforms approximates 1:1.  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