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.
The gene, designated hep, coding for a heparinase that degrades both heparin and heparan sulfate, was cloned from Bacillus circulans HpT298. Nucleotide sequence analysis showed that the open reading frame of the hep gene consists of 3,150 bp, encoding a precursor protein of 1,050 amino acids with a molecular mass of 116.5 kDa. A homology search found that the deduced amino acid sequence has partial similarity with enzymes belonging to the family of acidic polysaccharide lyases that degrade chondroitin sulfate and hyaluronic acid. Recombinant mature heparinase (111.2 kDa) was produced by the addition of IPTG from Escherichia coli harboring pETHEP with an open reading frame of the mature hep gene and was purified to homogeneity by SDS-polyacrylamide gel electrophoresis. Analyses of substrate specificity and degraded disaccharides indicated that the recombinant enzyme acts on both heparin and HS, as does heparinase purified from the wild-type strain.  相似文献   

3.
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   

4.
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   

5.
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   

6.
Infection of cells with Classical swine fever virus (CSFV) is mediated by the interaction of envelope glycoprotein E(rns) and E2 with the cell surface. In this report we studied the role of the cell surface glycoaminoglycans (GAGs), chondroitin sulfates A, B, and C (CS-A, -B, and -C), and heparan sulfate (HS) in the initial binding of CSFV strain Brescia to cells. Removal of HS from the surface of swine kidney cells (SK6) by heparinase I treatment almost completely abolished infection of these cells with virus that was extensively passaged in swine kidney cells before it was cloned (clone C1.1.1). Infection with C1.1.1 was inhibited completely by heparin (a GAG chemically related to HS but sulfated to a higher extent) and by dextran sulfate (an artificial highly sulfated polysaccharide), whereas HS and CS-A, -B, and -C were unable to inhibit infection. Bound C1.1.1 virus particles were released from the cell surface by treatment with heparin. Furthermore, C1.1.1 virus particles and CSFV E(rns) purified from insect cells bound to immobilized heparin, whereas purified CSFV E2 did not. These results indicate that initial binding of this virus clone is accomplished by the interaction of E(rns) with cell surface HS. In contrast, infection of SK6 cells with virus clones isolated from the blood of an infected pig and minimally passaged in SK6 cells was not affected by heparinase I treatment of cells and the addition of heparin to the medium. However, after one additional round of amplification in SK6 cells, infection with these virus clones was affected by heparinase I treatment and heparin. Sequence analysis of the E(rns) genes of these virus clones before and after amplification in SK6 cells showed that passage in SK6 cells resulted in a change of an Ser residue to an Arg residue in the C terminus of E(rns) (amino acid 476 in the polyprotein of CSFV). Replacement of the E(rns) gene of an infectious DNA copy of C1.1.1 with the E(rns) genes of these virus variants proved that acquisition of this Arg was sufficient to alter an HS-independent virus to a virus that uses HS as an E(rns) receptor.  相似文献   

7.
Heparinase (EC 4.2.2.7) isolated from Flavobacterium heparinum was purified to homogeneity by a combination of hydroxylapatite chromatography, repeated gel filtration chromatography, and chromatofocusing. Homogeneity was established by the presence of a single band on both sodium dodecyl sulfate and acid-urea gel electrophoretic systems. Amino acid analysis shows that the enzyme contains relatively high amounts of lysine residues (9%) consistent with its cationic nature (pI 8.5) but contains only 4 cysteine residues/polypeptide. The molecular weight of heparinase was estimated to be 42,900 +/- 1,000 daltons by gel filtration and 42,700 +/- 1,200 daltons by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme is very specific, acting only on heparin and heparan monosulfate out of 12 similar polysaccharide substrates tested. It has an activity maximum at pH 6.5 and 0.1 M NaCl and a stability maximum at pH 7.0 and 0.15 M NaCl. The Arrhenius activation energy was found to be 6.3 kcal/mol. However, the enzyme is very sensitive to thermal denaturation and loses activity very rapidly at temperatures over 40 degrees C. Kinetic studies of the heparinase reaction at 37 degrees C gave a Km of 8.04 X 10(-6) M and a Vm of 9.85 X 10(-5) M/min at a protein concentration of 0.5 microgram/ml. By adapting batch procedures of hydroxylapatite and QAE (quaternary aminoethyl)-Sephadex chromatography, gram quantities of heparinase that is nearly free of catalytic enzyme contaminants can be purified in 4-5 h.  相似文献   

8.
Heparinase I from Flavobacterium heparinum, a source of diverse polysaccharidases, suffers from low yields, insufficient purity for structural studies and insolubility when expressed as a recombinant product in Escherichia coli that is devoid of glycosaminoglycan polysaccharidases. In this study, cDNA coding for the orthologue of F. heparinum heparinase I was constructed from genomic information from the mammalian gut symbiont Bacteroides thetaiotaomicron and expressed in E. coli as a fusion protein with GST at the N-terminus. This resulted in high yield (30 mg/g dry bacteria) of soluble product and facilitated one-step affinity purification to homogeneity. Purified heparinase I bearing the GST fusion exhibited a Km of 2.3 μM and Vmax of 42.7 μmol/min with a specific activity of 164 U/mg with heparin (average 12,000 Da) as substrate. The results indicate a 2-fold improvement in yield, specific activity and affinity for heparin as substrate over previous reports. The data suggest that the heparinase I from the gut symbiont exhibits a higher intrinsic affinity for heparin than that from F. heparinum. The purified GST fusion enzyme exhibited a requirement for Ca2+ and a pH optimum between 6.7 and 7.3 that was similar to the enzyme freed of the N-terminal GST portion. Our study revealed that catalytic activity of heparinase I requires a reducing environment. The GST facilitated immobilization of heparinase I in solid phase either for clinical purposes or for structural studies in absence of interference by contaminating polysaccharidases.  相似文献   

9.
Previous studies have suggested that the attachment of bovine herpesvirus 1 (BHV-1) to permissive cells is mediated by its major glycoproteins B (gB), C (gC), and D (gD). In order to gain further insight into the mechanism of the BHV-1 attachment process, we purified authentic gB, gC, and gD from BHV-1-infected cells and membrane anchor-truncated, soluble gB, gC, and gD from stably transfected cell lines by affinity chromatography and examined their cell-binding properties on Madin-Darby bovine kidney cells. All of the glycoproteins tested exhibited saturable binding to Madin-Darby bovine kidney cells. All of the glycoproteins tested exhibited saturable binding to Madin-Darby bovine kidney cells. Addition of exogenous heparin or treatment of cells with heparinase to remove cellular heparan sulfate (HS) prevented both gC and gB from binding to cells but had no effect on gD binding. An assessment of competition between gB, gC, and gD for cell binding revealed that gC was able to inhibit gB binding, whereas other combinations showed no effect. Cell-bound gC could be dissociated by heparin or heparinase treatment. The response of bound gB to heparin and heparinase treatments differed for the authentic and soluble forms; while soluble gB was susceptible to the treatment, a significant portion of cell-bound authentic gB was resistant to the treatment. Binding affinity analysis showed that soluble gB and both forms of gC and gD each had single binding kinetics with comparable dissociation constants (Kds), ranging from 1.5 x 10(-7) to 5.1 x 10(-7) M, whereas authentic gB exhibited dual binding kinetics with Kd1 = 5.2 x 10(-7) M and Kd2 = 4.1 x 10(-9) M. These results demonstrate that BHV-1 gC binds only to cellular HS, gD binds to a non-HS component, and gB initially binds to HS and then binds with high affinity to a non-HS receptor. Furthermore, we found that while authentic gB was able to inhibit viral plaque formation, soluble gB, which retains the HS-binding property but lacks the high-affinity binding property, was defective in this respect. These results suggest that the interaction between gB and its high-affinity receptor may play a critical role in the virus entry process.  相似文献   

10.
Most, although not all, samples of commercial calf thymus DNA were strongly inhibitory to DNA polymerase alpha; the inhibition made the DNA useless as a template for this enzyme. In a pre-assembled DNA polymerase assay mixture (minus enzyme but including activated DNA) the inhibition tended to diminish with time but at a rate that was not predictable, and some inhibition usually persisted. It was concluded that the inhibition was the result of contamination of the DNA by a heparin-like material on the basis of the following: 1) the inhibition could be reversed by treatment of the DNA with heparinase; 2) both the endogenous inhibitory effect of calf thymus DNA as well as the inhibitory effect of heparin on DNA polymerase alpha are reversed by protamine (which is known to prevent the antithrombin activity of heparin); 3) both the endogenous inhibition and inhibition by heparin are also reversed by ampholyte (which also prevents the antithrombin activity of heparin); and 4) both the endogenous and the heparin-induced inhibitory effects display the same spectrum of activity against mammalian DNA polymerases, i.e. both DNA polymerases alpha and delta are extremely sensitive whereas, DNA polymerases beta and gamma are resistant. The last result also suggests the use of heparin as a specific inhibitor of purified mammalian DNA polymerases alpha and delta, similar to the use of aphidicolin.  相似文献   

11.
Chen J  Duncan MB  Carrick K  Pope RM  Liu J 《Glycobiology》2003,13(11):785-794
Heparan sulfate 3-O-sulfotransferase transfers sulfate to the 3-OH position of a glucosamine to generate 3-O-sulfated heparan sulfate (HS), which is a rare component in HS from natural sources. We previously reported that 3-O- sulfotransferase isoform 5 (3-OST-5) generates both an antithrombin-binding site to exhibit anticoagulant activity and a binding site for herpes simplex virus 1 glycoprotein D to serve as an entry receptor for herpes simplex virus. In this study, we characterize the substrate specificity of 3-OST-5 using the purified enzyme. The enzyme was expressed in insect cells using the baculovirus expression approach and was purified by using heparin-Sepharose and 3',5'-ADP- agarose chromatographies. As expected, the purified enzyme generates both an antithrombin binding site and a glycoprotein D binding site. We isolated IdoUA-AnMan3S and IdoUA-AnMan3S6S from nitrous acid-degraded 3-OST-5-modified HS (pH 1.5), suggesting that 3-OST-5 enzyme sulfates the glucosamine residue that is linked to an iduronic acid residue at the nonreducing end. We also isolated a disaccharide with a structure of DeltaUA2S-GlcNS3S and a tetrasaccharide with a structure of DeltaUA2S-GlcNS-IdoUA2S-GlcNH23S6S from heparin lyases-digested 3-OST-5-modified HS. Our results suggest that 3-OST-5 enzyme sulfates both N-sulfated glucosamine and N-unsubstituted glucosamine residues. Taken together, the results indicate that 3-OST-5 has broader substrate specificity than those of 3-OST-1 and 3-OST-3. The unique substrate specificity of 3-OST-5 serves as an additional tool to study the mechanism for the biosynthesis of biologically active HS.  相似文献   

12.
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   

13.
Identification with specific heparan sulfate-lyases, heparitinase I and heparinase of the constitutional unsaturated disaacharide (ΔDi-SHS) derived from heparan sulfate (HS) isomers and heparin was achieved using high-performance liquid chromatography (HPLC) with a sulfonated styrene-divinylbenzene copolymer. Eight ΔDi-SHS products derived from HS isomers were identified. Enzymatic digestion with heparitinase I and heparinase converts heterogeneous sulfated HS isomers and heparin into different ΔDi-SHS. The practical application of these enzymes was examined using specific enzymes and HPLC. In a patient with Hurler syndrome, eight individual Δi-SHS were identified in urinary HS isomers.  相似文献   

14.
Teesalu K  Panarina M  Uibo O  Uibo R  Utt M 《Amino acids》2012,42(2-3):1055-1064
Autoantibodies from patients with celiac disease (CD) can influence transglutaminase 2 (TG2) activity and its cellular functions, but the exact mechanisms have remained unknown. Our objective was to study whether autoantibodies could modulate TG2 binding to heparin/heparan sulfate (HS) and intestinal epithelial cell attachment to fibronectin-TG2 matrix. Anti-TG2 antibodies were purified by TG2 affinity chromatography from sera of patients with active CD. Serum and antibody effects on TG2 binding to heparin/HS, on transamidase activity of TG2, as well as on Caco-2 cell attachment to fibronectin-TG2 matrix were assessed using microplate assays. Both sera and purified anti-TG2 antibodies from CD patients with high anti-TG2 IgA levels reduced TG2 binding to heparin/HS as compared with those with low anti-TG2 IgA or controls. There was a negative correlation between anti-TG2 IgA levels and TG2 binding to heparin/HS. Treatment of fibronectin-TG2 coated wells with CD patients' sera or purified anti-TG2 antibodies reduced attachment of Caco-2 cells onto the plate as compared with the control samples. The effect of CD patients' antibodies on Caco-2 cell attachment to fibronectin-TG2 matrix occurred independently of the inhibition of cell adhesion by Arg-Gly-Asp sequence containing peptides. Anti-TG2 autoantibodies had no effect on transamidase activity of TG2 in vitro. We suggest that modulation of adhesion function of TG2 by autoantibodies from patients with CD could be related to the inhibition of TG2 binding to HS residues of cell surface proteoglycans and could have possible implications for CD pathogenesis.  相似文献   

15.
An immobilized enzyme reactor has been developed to remove heparin, the anticoagulant that is required in all extracorporeal devices for patients undergoing open-heart surgery or kidney dialysis. The device uses the enzyme heparinase (EC 4.2.2.7), which is covalently linked to agarose with cyanogen bromide. A critical parameter in the development of a model for the degradation of heparin catalyzed by immobilized heparinase is the radial concentration profile of the enzyme within the agarose matrix. Experimental determinations of bound enzyme con centrations have been conducted previously for several enzyme systems using radioactive or fluorescent labels. For the development of the heparinase reactor it is necessary to use catalytically but not electrophoretically pure enzyme, and thus it is not possible to use the labeling techniques. To obtain information about the bound enzyme distribution, an experimental study of the intrinsic binding kinetics of heparinase to cyanogen bromide-activated agarose was conducted. The binding reaction was studied as a function of both the concentration of heparinase and the gel-reactive group. At conditions of functional group excess, the binding kinetics were pseudo first order in heparinase concentration with a rate constant equal to 0.12 C(c[triple chemical bond]n) (h(-1)), where C(c[triple chemical bond]n) is the gel-reactive group concentration. The reactive group concentration remained constant within the 2-4-h experiments. Competitive binding between heparinase and the protein contaminants was unimportant. A model was formulated for the immobilization procedure based on the diffusion of heparinase within the porous network and the binding kinetics as determined above. The model predicted the immobilization of heparinase to be kinetically controlled and the enzyme to distribute uniformly within the agarose matrix. These experimental techniques could be applied to predict the immobilized enzyme distribution for different enzyme systems that are not electrophoretically pure.  相似文献   

16.
Polyclonal IgG rabbit antibodies were prepared against a purified heparinase from Flavobacterium heparinum. Immuno-affinity purification of crude and partially purified heparinase is described. The resulting enzyme was of comparable purity to that prepared using the standard multistep purification scheme. The antibodies prepared were found to increase the activity of bound heparinase.  相似文献   

17.
Recombinant heparinase III (rHepIII) from Bacteroides stercoris HJ-15 was cloned, expressed, and characterized. The full-length heparinase III gene from B. stercoris HJ-15 was identified by Southern blotting, and the sequence was deposited in GenBank. The heparinase III gene, which is 2,001-bp long, was cloned and overexpressed in Escherichia coli; highly active rHepIII was easily purified using only one step of immobilized Ni2+ affinity column chromatography. Enzymatic properties and substrate specificities of rHepIII were assessed, and its kinetic constants were calculated. rHepIII was most active in 50 mM sodium phosphate buffer with 350 mM NaCl (pH 6.6) at 45°C. Through amino acid modification studies and site-directed mutagenesis assay, cysteines and histidines were identified as crucial residues for enzymatic activity. Moreover, this enzyme digested not only heparan sulfate but also heparin and hyaluronic acid, and their degradation products were verified by strong anion exchange/high-performance liquid chromatography. These characteristics, including active residues and substrate specificities were interesting compared with those of existing heparinase III from other species. We anticipate that the convenience of purification and the characteristics of this enzyme will make it a powerful tool for studies of glycosaminoglycans and their lyases.  相似文献   

18.
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   

19.
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   

20.
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   

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