首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A type of heparinase (heparin lysase, no EC number) was isolated from the periplasmic space of a novel species of Sphingobacterium by three-step osmotic shock. It was further purified to apparent homogeneity by a combination of SP-sepharose and Source 30S chromatographies with a final specific activity of 17.6 IU/mg protein and purification factor of 13-fold. MALDI-TOF mass spectrum of the purified heparinase gave a molecular mass of 75,674 Da of the native enzyme. Peptide mass spectrum showed poor homogeneity with the database in the peptide bank. Inhibition of the enzyme activity by N-acetylimidazole indicated that tyrosine residues were necessary for enzyme activity. K(m) and V(max) of the heparinase for de-o-sulfated-N-acetyl heparin were 42 micro M and 166 microM/min/mg protein, respectively. The heparinase showed similar activity on both heparin and heparan sulfate, except for the heparin from bovine lung. The heparinase exhibited only 8.3% of the activity when de-N-sulfated heparin was used as the substrate, but N-acetylation of the de-N-sulfated heparin restored the activity to 78.4%. Thus modification of N-site in heparin structure was favorable for heparinase activity. On the other hand, de-o-sulfation in heparin showed positive effects on the heparinase activity, since the enzyme activity for N-acetyl-de-o-sulfated heparin was increased by 150%. Based on the present findings, the sphingobacterial heparinase differed from flavobacterial and other reported heparinases in molecular mass, composition, charge properties, active site, substrate specificities and other important characteristics, suggesting that it a novel heparin lysase distinct from those from other sources.  相似文献   

2.
Heparinase production by Flavobacterium heparinum.   总被引:11,自引:3,他引:8       下载免费PDF全文
Heparinase production by Flavobacterium heparinum in complex protein digest medium, with heparin employed as the inducer, has been studied and improved. The maximum productivity of heparinase has been increased 156-fold over that achieved by previously published methods to 375 U/liter per h in the complex medium. Rapid deactivation of heparinase activity, both specific and total, was observed at the onset of the stationary phase. Nutritional studies on growth and heparinase production showed an obligate requirement for L-histidine and no vitamin requirement. L-Methionine partially relieved the L-histidine requirement. A defined medium containing glucose, ammonium sulfate, basal salts, L-methionine, and L-histidine was developed for growth and heparinase production. The growth rate in this medium was 0.21 h-1, which is 40%, higher than that in complex medium. The maximum volumetric productivity of heparinase in the defined medium was increased to 1,475 U/liter per h, providing a 640-fold increase over that achieved by previously published methods. No rapid deactivation was observed. An examination of alternate inducers for heparinase showed that heparin degradation products, hyaluronic acid, heparin monosulfate, N-acetyl-D-glucosamine, and maltose, induce heparinase in complex medium. An Azure A assay was modified and fully developed to measure the heparin concentration during fermentation and the heparinase specific activity of crude extracts of F. heparinum obtained from sonication, thus negating the need for further purification to measure activity."  相似文献   

3.
A rapid, sensitive, assay for enzymes that degrade heparin is described. The procedure is based on the interference of heparin with color development during the interaction of protein with the dye Coomassie brilliant blue. The loss of this property when the glycosaminoglycan is degraded by heparinase can be used to quantify activity of the enzyme in pure form, or in complex biological samples such as tissue homogenates or serum. The assay is also suitable for studying dependence of heparinase activity under conditions such as varying pH and temperature.  相似文献   

4.
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   

5.
Purification and characterization of a novel heparinase   总被引:2,自引:0,他引:2  
A unique heparinase was isolated from a recently discovered Gram-negative soil bacterium. The enzyme (heparinase III) was purified by hydroxylapatite chromatography, chromatofocusing, and gel permeation chromatography. The enrichment was 48x, and the specific activity of catalytically pure heparinase was 127 IU/mg of protein. Similar to the heparinase I from Flavobacterium heparinum, heparinase III also degrades heparin to mainly disaccharide fragments. It is specific for heparin and also breaks down heparan sulfate, but not hyaluronic acid and chondroitin sulfate. Heparinase III, however, differs markedly from heparinase I in several other aspects: it has a higher molecular mass (94 versus 43 kDa), pI (9.2 versus 8.5), its Km and kcat are different, and it has a higher energy of activation (15.6 versus 6.3 kcal/mol). Optimal activity was also found at higher pH (7.6 versus 6.5) and temperature (45 versus 37 degrees C). Furthermore, the amino acid composition of heparinase III is quite different from that of heparinase I.  相似文献   

6.
A novel type of heparinase (heparin lyase, no EC number) has been purified from Bacteroides stercoris HJ-15, isolated from human intestine, which produces three kinds of heparinases. The enzyme was purified to apparent homogeneity by a combination of QAE-cellulose, DEAE-cellulose, CM-Sephadex C-50, hydroxyapatite, and HiTrap SP chromatographies with a final specific activity of 19.5 mmol/min/mg. It showed optimal activity at pH 7.2 and 45 degrees C and the presence of 300 mM KCl greatly enhanced its activity. The purified enzyme activity was inhibited by Cu(2+), Pb(2+), and some agents that modify histidine and cysteine residues, and activated by reducing agents such as dithiothreitol and 2-mercaptoethanol. This purified Bacteroides heparinase is an eliminase that shows its greatest activity on bovine intestinal heparan sulfate, and to a lesser extent on porcine intestinal heparan sulfate and heparin. This enzyme does not act on acharan sulfate but de-O-sulfated acharan sulfate and N-sulfoacharan sulfate were found to be poor substrates. The substrate specificity of this enzyme is similar to that of Flavobacterial heparinase II. However, an internal amino acid sequence of the purified Bacteroides heparinase shows significant (73%) homology to Flavobacterial heparinase III and only 43% homology to Flavobacterial heparinase II. These findings suggest that the Bacteroidal heparinase is a novel enzyme degrading GAGs.  相似文献   

7.
Heparin lyase I (heparinase I) specifically depolymerizes heparin, cleaving the glycosidic linkage next to iduronic acid. Here, we show the crystal structures of heparinase I from Bacteroides thetaiotaomicron at various stages of the reaction with heparin oligosaccharides before and just after cleavage and product disaccharide. The heparinase I structure is comprised of a β-jellyroll domain harboring a long and deep substrate binding groove and an unusual thumb-resembling extension. This thumb, decorated with many basic residues, is of particular importance in activity especially on short heparin oligosaccharides. Unexpected structural similarity of the active site to that of heparinase II with an (α/α)6 fold is observed. Mutational studies and kinetic analysis of this enzyme provide insights into the catalytic mechanism, the substrate recognition, and processivity.  相似文献   

8.
Ahn SC  Kim BY  Oh WK  Park YM  Kim HM  Ahn JS 《Life sciences》2006,79(17):1661-1665
Heparanase has been previously associated with the metastatic potential, inflammation, and angiogenesis of tumor cells. Heparanase activity has been detected by means of UV absorption, radiolabeled substrates, electrophoretic migration, and heparan sulfate affinity assays. However, those methods have proven to be somewhat problematic with regards to application to actual biological samples, the accessibility of the immobilized substrates, experimental sensitivity, and the separation of degraded products. Rather than focusing on heparanase activity, then, we have developed a rapid, alternative colorimetric heparinase assay, on the basis of the recent finding that sulfated disaccharides generated from heparin by bacterial heparinase exhibit biological properties comparable to those from heparan sulfate by mammalian heparanase. In this study, the concentrations of porcine heparin and bacterial heparinase I were determined using a Sigma Diagnostics Kit. Morus alba was selected as a candidate through this assay system, and an inhibitor, resveratrol, was purified from its methanol extract. Its anti-metastatic effects on the pulmonary metastasis of murine B16 melanoma cells were also evaluated. Our findings suggest that this assay may prove useful as a diagnostic tool for heparinase inhibition, as an alternative anti-metastatic target.  相似文献   

9.
Activation of heparin cofactor II by heparin oligosaccharides   总被引:1,自引:0,他引:1  
Heparin was partially depolymerized with heparinase or nitrous acid. The resulting oligosaccharides were fractionated by gel filtration chromatography and tested for the ability to stimulate inhibition of thrombin by purified heparin cofactor II or antithrombin. Oligosaccharides containing greater than or equal to 18 monosaccharide units were active with antithrombin, while larger oligosaccharides were required for activity with heparin cofactor II. Intact heparin molecules fractionated on a column of immobilized antithrombin were also tested for activity with both inhibitors. The relative specific activities of the unbound heparin molecules were 0.06 with antithrombin and 0.76 with heparin cofactor II in comparison to unfractionated heparin (specific activity = 1.00). We conclude that heparin molecules much greater than 18 monosaccharide units in length are required for activity with heparin cofactor II and that the high-affinity antithrombin-binding structure of heparin is not required.  相似文献   

10.
The effect of various carbon, nitrogen and sulfur sources on the production of heparinase by Flavobacterium heparinum in defined medium in the presence and absence of heparin as the inducer has been studied. Carbon catabolite repression has been observed in defined medium containing one of several carbon sources including simple sugars, alcohols and organic acids. Fed batch fermentations result in 10 g/l of cells and heparinase titers as high as 100,000 U/l by avoiding carbon catabolite repression. Growth on heparin as a sole carbon source resulted in both a high growth rate of 0.12 h–1 and a high specific activity of 18 U/mg. Specific heparinase activity was markedly reduced when the end products of heparin catabolism were used as carbon, nitrogen or sulfur sources in defined medium. In defined medium with a low sulfate concentration, of less than 10–3 M, specific activities as high as 8 U/mg have been observed even in the absence of the normally required inducer, heparin.  相似文献   

11.
The activity of tissue plasminogen activator (t-PA) and urokinase-type plasminogen activator (u-PA) is stimulated by heparin. Heparin binds tightly to t-PA, u-PA, and plasminogen and decreases the usual stimulatory effect of fibrin on t-PA activity. In the present study we have found that low molecular weight heparin (LMW-heparin) preparations obtained by nitrous acid depolymerization or heparinase treatment of standard heparin have different properties with respect to their interaction with the fibrinolytic system. LMW-heparin prepared by either method does not stimulate plasmin formation by t-PA. However, these preparations of heparin still efficiently accelerate the inhibition of thrombin by antithrombin III. Binding data show that LMW-heparin does not bind t-PA and Glu-plasminogen and only binds very weakly to Lys-plasminogen. These results illustrate that it is possible to selectively destroy the fibrinolytic stimulating properties of heparin while leaving the classical anticoagulant characteristics intact.  相似文献   

12.
Immobilization of biologically active proteins is of great importance to research and industry. Cellulose is an attractive matrix and cellulose-binding domain (CBD) an excellent affinity tag protein for the purification and immobilization of many of these proteins. We constructed two vectors to enable the cloning and expression of proteins fused to the N- or C-terminus of CBD. Their usefulness was demonstrated by fusing the heparin-degrading protein heparinase I to CBD (CBD-HepI and HepI-CBD). The fusion proteins were over-expressed in Escherichia coli under the control of a T7 promoter and found to accumulate in inclusion bodies. The inclusion bodies were recovered by centrifugation, the proteins were refolded and recovered on a cellulose column. The bifunctional fusion protein retained its abilities to bind to cellulose and degrade heparin. C-terminal fusion of heparinase I to CBD was somewhat superior to N-terminal fusion: Although specific activities in solution were comparable, the latter exhibited impaired binding capacity to cellulose. CBD-HepI-cellulose bioreactor was operated continuously and degraded heparin for over 40 h without any significant loss of activity. By varying the flow rate, the mean molecular weight of the heparin oligosaccharide produced could be controlled. The molecular weight distribution profiles, obtained from heparin depolymerization by free heparinase I, free CBD-HepI, and cellulose-immobilized CBD-HepI, were compared. The profiles obtained by free heparinase I and CBD-HepI were indistinguishable, however, immobilized CBD-HepI produced much lower molecular weight fragments at the same percentage of depolymerization. Thus, CBD can be used for the efficient production of bioreactors, combining purification and immobilization into essentially a single step.  相似文献   

13.
Glycosaminoglycans are known to participate in the attachment of several chlamydial strains. We studied the effect of heparin, enoxaparin, low-molecular-weight heparin, chondroitin sulfate A, and heparinase I on the infectivity of Chlamydia pneumoniae strain CWL029 and two Finnish isolates, Kajaani 7 and Parola, in an HL cell line which is epithelial in origin. Two Chlamydia trachomatis strains, L2 and E, were used for comparison. The infectivity of all C. pneumoniae strains and C. trachomatis serovar E was inhibited not only by heparin derivatives but also by chondroitin sulfate A and heparinase treatment. Treatment of host cells with heparin derivatives and heparinase was also inhibitory. Different chlamydial strains and species seem, however, to vary in their ability to use heparin in their attachment to host cells.  相似文献   

14.
A heparinase that degrades both heparin and heparan sulfate (HS) was purified to homogeneity from the cell-free extract of Bacillus circulans HpT298. The purified enzyme had a single band on SDS-polyacrylamide gel electrophoresis with an estimated molecular mass of 111,000. The enzyme showed optimal activity at pH 7.5 and 45 degrees C, and its activity was stimulated in the presence of 5 mM CaCl2, BaCl2, or MgCl2. Analysis of substrate specificity and degraded disaccharides demonstrated that the enzyme acts on both heparin and HS, similar to heparinase II from Flavobacterium heparinum.  相似文献   

15.
Because tumors and other diseases are characterized by increased heparanase levels, human heparanase is a promising drug target and diagnostic marker. Therefore, methods are needed to determine heparanase activity and to examine potential inhibitors. Because of substrate comparability, we used the bacterial enzyme heparinase II (heparinase) for the assay development. Usually the substrate of heparanase assays is heparan sulfate, which has several disadvantages. Because of that, we used fondaparinux, which is being cleaved by both heparanase and heparinase. Two concepts to detect its degradation were examined: measurement of anti-factor Xa activity of fondaparinux and its direct quantification with the fluorescent sensor polymer-H. Using fondaparinux as substrate, the anti-factor Xa assay was shsown to be appropriate to determine heparinase activity. The detection with polymer-H was easier and even faster to perform. Linearity was given with fondaparinux as well as heparan sulfate, and heparin as substrates, but fondaparinux turned out to be most suitable. By modifications (incubation time, fondaparinux concentration, and polymer-H concentration), the limit of quantification and the linear range can be adapted to the respective requirements. In conclusion, a simple, accurate, and robust heparinase assay was developed. It is suitable for heparinase quality control and testing heparinase inhibitors and could be adapted to heparanase.  相似文献   

16.
Five chemically modified heparins were derived from native pig mucosal heparin (pig heparin Is). These were de-N-sulphated heparin (heparin IH), N-acetylheparin (heparin IA), de-N/O-sulphated heparin (heparin IVH), de-O-sulphated heparin (heparin IVs) and de-O-sulphated N-acetyl-heparin (heparin IVA). Their structures were studied by 13C-NMR spectroscopy at 90.56 MHz. Native heparin and the derivatives were incubated with Flavobacterium heparinase II at 25 degrees C. The progress of degradation was followed by the delta A235 and the final composition examined by gel filtration with Bio-Gel P-4. Native heparin (Is) was readily degraded by heparinase II and, with the exception of heparin IVH for which degradation was negligible, the chemically modified derivatives were also degraded. Approximately 90% of the saccharides from heparins Is, IA, IVs and IVA were disaccharides and tetrasaccharides. For heparin IH, which was degraded more slowly, the proportion was 65%. Heparins Is, IVs and IVA underwent initial rapid degradation. The digestion of heparin Ia proceeded rapidly after an initial lag phase. The undegraded polymers produced similar elution profiles from Bio-Gel P-4. Following the action of heparinase II on heparins Is, IA, IVs and IVA, the elution profiles revealed a major peak of disaccharides and minor peaks of higher oligomers. The profile of heparin IH revealed a greater proportion of intermediate-molecular-mass saccharides. Our results demonstrate a broad specificity for heparinase II. It is capable of lysing both N-acetylated and N-sulphated heparins independent of O-sulphation. Heparinase II will also degrade heparin derivatives that are non-N-substituted provided that they are O-sulphated.  相似文献   

17.
Sulfur regulation of heparinase and sulfatases in Flavobacterium heparinum   总被引:2,自引:0,他引:2  
Sulfur regulation of heparinase synthesis and sulfatase synthesis was studied in Flavobacterium heparinum. Heparinase synthesis was strongly repressed by sulfate and L-cysteine, while the activity of this enzyme showed little or no inhibition by these compounds. Heparinase was synthesized in the absence of heparin when L-methionine was used as the sole sulfur source. The sulfatases produced by F. heparinum, which include the sulfatases involved in heparin catabolism, were also studied. At least some of the sulfatase activity was regulated by sulfur compounds in a manner similar to heparinase regulation. L-Cysteic acid and taurine were not suitable sulfur sources to support the growth of F. heparinum.  相似文献   

18.
Sulfur regulation of heparinase synthesis and sulfatase synthesis was studied in Flavobacterium heparinum. Heparinase synthesis was strongly repressed by sulfate and L-cysteine, while the activity of this enzyme showed little or no inhibition by these compounds. Heparinase was synthesized in the absence of heparin when L-methionine was used as the sole sulfur source. The sulfatases produced by F. heparinum, which include the sulfatases involved in heparin catabolism, were also studied. At least some of the sulfatase activity was regulated by sulfur compounds in a manner similar to heparinase regulation. L-Cysteic acid and taurine were not suitable sulfur sources to support the growth of F. heparinum.  相似文献   

19.
Heparinase and heparitinase were separated from an extract of Flavobacterium heparinum, induced with heparin by using column chromatography on hydroxylapatite. As the heparinase preparation contained chondroitinases B and C, chondroitinase B was removed by rechromatography on a hydroxylapatite column. Chondroitinase C was then eliminated by column chromatography on O-phosphono(“phospho”)-cellulose. The heparinase preparation thus obtained was free from sulfoamidase for 2-deoxy-2-sulfoamino-D-glucose (GlcN-2S), sulfatase for 2-amino-2-deoxy-6-O-sulfo D-glucose (GlcN-6S), as well as (δ4,5glycosiduronase for the unsaturated disaccharides obtained from heparin. The remaining sulfatase for 4-deoxy-α-L-thero-hex-4-enopyranosyluronic acid 2-sulfate (δUA-2S) in the heparinase preparation was removed by affinity chromatography with dermatan sulfate-bound AH-Sepharose 4B coated with dermatan sulfate. The heparitinase preparation separated by column chromatography on hydroxyla patite was purified by affinity chromatography with heparin-bound AH-Sepharose 4B coated with heparin. Sulfatase for 2-amino-2-deoxy-6-O-sulfo-D-glucose (GlcN-6S) and δ4,5glycosiduronase for the unsaturated disaccharides obtained from heparin were removed by this chromatography. Sulfatase for 4-deoxy-α-L-threo-hex-4-enopyranosyluronic acid 2-sulfate (δUA-2S) remaining in the heparitinase preparation was finally removed by column chromatography on hydroxylapatite. The recoveries of the purified preparations of heparinase and heparitinase were estimated to be 39 and 50%, respectively, from the crude enzyme fractions obtained by the first column chromatography on hydroxyl- patite. The purified heparinase and heparitinase were free from all enzymes that could degrade the sulfated unsaturated disaccharides produced from heparin with heparinase.  相似文献   

20.
The development of a safe and efficient bioreactor design has remained a challenge for the clinical application of immobilized enzymes. Specifically, the use of immobilized heparinase I has been the target of many studies to make heparin anticoagulation therapy safer for the critically ill patient with kidney failure or heart disease. We have investigated the use of Taylor-Couette flow for a novel type of bioreactor. In a previous study, we showed that the fluidization of agarose immobilized heparinase within Taylor vortices in whole blood can lead to extensive blood damage in the form of cell depletion and hemolysis. Based on these findings, we designed and developed a reactor, referred to as vortex-flow plasmapheretic reactor (VFPR), that incorporated plasmapheresis and fluidization of the agarose in the reactive compartment, separate from the whole-blood path. In the present study, immobilized heparinase I was tested as a means of achieving regional heparinization of a closed circuit. This is a method in which heparin is infused into the extracorporeal circuit predialyzer and neutralized postdialyzer. Saline studies were performed with an immobilized heparinase I-packed bed and with the VFPR. An in vitro feasibility study was performed with the VFPR using human blood. The VFPR achieved heparin conversions of 44 +/- 0.5% and 34 +/- 2% in saline and blood, respectively. In addition, the VFPR caused no blood damage. We report a novel method to achieve fluidization which depended on secondary, circumferencial flow, and was independent of the primary flow through the device.  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