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.
Tissue plasminogen activator (t-PA) is an important initiator of fibrinolysis. The t-PA polypeptide has four potential N-glycosylation sites of which three are occupied in type I (Asn-117, -184, and -448) and two in type II (Asn-117 and -448). In an effort to elucidate the factors controlling the expression of N-linked oligosaccharides on this polypeptide, we have used a combination of sequential exoglycosidase digestion, methylation analysis, and controlled acetolysis to determine the oligosaccharide structures at each of the N-glycosylation sites of type I and type II t-PA when isolated from a human colon fibroblast cell strain and from a Bowes melanoma cell line. Our results suggest the following: (i) type I and type II t-PA are N-glycosylated in an identical way at Asn-117 and Asn-448, when isolated from the same cell line; (ii) Asn-117 is predominantly associated with oligomannose-type structures in all cases; (iii) Asn-184 and Asn-448 are predominantly associated with complex-type structures when t-PA is isolated from fibroblast cells, but with both complex- and oligomannose-type structures when isolated from melanoma cells; (iv) fibroblast cell derived t-PA is associated with both neutral and sialylated oligosaccharides, while melanoma cell derived t-PA is also associated with sulfated oligosaccharides, which are located exclusively at Asn-448 of type II t-PA; (v) no complex-type structures occur in common between t-PA from the two cell lines. These results indicate that the t-PA glycoprotein is secreted by each cell line as a set of glycoforms, each glycoform being unique with respect to the nature and disposition of oligosaccharides on a common polypeptide. Further, the two cell lines express no glycoform in common, despite expressing the same t-PA polypeptide. The implications of these results for both the control of oligosaccharide processing in different cell lines and the genetic engineering of mammalian glycoproteins are discussed.  相似文献   

2.
Kinetic parameters (k(Pg) and K(Pg)) were determined for activation of Glu-plasminogen (Glu-Pg) and Lys-plasminogen (Lys-Pg) type I (with N-linked carbohydrate chain at Asn-289) and type II (with unsubstituted Asn-289) by plasmin-staphylokinase (Pm-STA) complex. The K(Pg) values for Glu-Pg I and Lys-Pg I (17.1 and 11.2 microM, respectively) were higher than those for Glu-Pg II and Lys-Pg II (14.9 and 5.4 microM, respectively), while only minor differences in the k(Pg) values were observed between plasminogens type I and type II. Soluble fibrin significantly increased the k(Pg)/K(Pg) values for activation of all four plasminogens due to a decrease in the K(Pg) values but did not alter the k(Pg) values. However, the activation of plasminogens type I was stimulated by fibrin lesser degree than that of plasminogens type II. These findings indicate that N-glycosylation of kringle 3 of plasminogen decreases the stability of Pm-STA-Pg ternary enzyme-substrate complex in solution as well as interferes with its formation and rearrangement on the fibrin surface.  相似文献   

3.
The effect of altering oligosaccharide structures at sites 184 and 448 of tissue plasminogen activator (tPA) has been examined. Alteration to high-mannose forms at sites 184 and 448 was accomplished by the growth of cells in the presence of deoxymannojirimycin (dMM). Modification to neutral, unsialylated forms at these sites was achieved by neuraminidase treatment of control preparations of tPA. Oligosaccharides at site 117 were not markedly affected by either treatment because structures at this site are high-mannose and not sialylated in untreated preparations. The effect on enzymatic activity and on a related property, lysine affinity, was determined. dMM treatment was found to increase both the lysine affinity and catalytic activity of tPA. Neuraminidase treatment increased enzyme activity, but was without effect on affinity for lysine. To evaluate the effects of alterations at site 184 and site 448, the catalytic activity and lysine affinity of type I and type II tPA were monitored individually. In the dMM-treated sample, type I tPA (with sugars at sites 117, 184 and 448) was found to have 2- to 3-fold increased catalytic activity and an affinity for lysine which was greater than that of type I from untreated preparations, but less than that of control type II tPA (containing sugar only at sites 117 and 448). In neuraminidase-treated type I, catalytic activity was also enhanced but lysine affinity remained unchanged. Type II from dMM- and neuraminidase-treated preparations had catalytic activity that was increased approximately 1.5-fold compared to untreated controls, whereas affinity for lysine was unchanged.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
The kinetics of activation of Glu-plasminogen (Glu-Pg) and Lys77-Pg by two-chain recombinant tissue plasminogen activator (t-PA) were determined in the presence of isolated protein components of the extracellular matrix (ECM) and compared to activation in the presence of fibrinogen and fibrinogen fragments and in the absence of added protein. Several ECM protein components were as effective as fibrinogen fragments at stimulating Pg activation. Stimulation of Glu-Pg activation resulted from both a decrease in Km and an increase in Vmax, whereas stimulation of Lys77-Pg was due primarily to increases in Vmax. The most effective stimulators of activation were basement membrane type IV collagen and gelatin which resulted in a 21- and 55-fold increase, respectively, in the kcat/Km of Glu-Pg (relative to a 10-fold increase observed with fibrinogen fragments). Amidolytic activity of t-PA was also enhanced up to 12-fold by ECM proteins. However, plasmin amidolytic activity was unaffected by the presence of added proteins. These data suggest that several ECM-associated proteins can enhanced the activation of Pg in the absence of fibrin.  相似文献   

5.
The role of Asn-linked oligosaccharide in the functional properties of both human tissue-type plasminogen activator (t-PA) and a genetic variant of t-PA was studied. Nonglycosylated and glycosylated wild-type t-PA were produced in mammalian cells which express recombinant t-PA. These proteins were compared in fibrin binding and 125I-labeled fibrin clot lysis assays, using purified components. The nonglycosylated form showed higher fibrin binding, as well as higher fibrinolytic potency than the glycosylated form. Subsequently, prevention of glycosylation of a t-PA variant which lacked the finger and epidermal growth factor domains (delta FE), was carried out in an attempt to enhance its fibrinolytic activity. Glycosylation was prevented by changing Asn to Gln; at Asn-117 to produce delta FE1X t-PA, and at Asn-117, -184, and -448 to produce delta FE3X t-PA. All variants were similar to wild-type t-PA in their catalytic dependence on fibrinogen fragments, fibrinolytic activity in fibrin autography analysis, and plasminogen activator activity. In a clot lysis assay, using citrated human plasma, the fibrinolytic potency of the variants were comparable to that of wild-type t-PA at activator concentrations of 17-51 nM (approximately 1-3 micrograms/ml). At 0.5-5.1 nM (approximately 0.03-0.3 micrograms/ml), however, the variant proteins had lower fibrinolytic potency than wild-type t-PA. Fifty percent lysis in 1.5 h for wild-type, delta FE, delta FE1X, and delta FE3X t-PA, required 2.5, 10, 7.5, and 5.5 nM t-PA, respectively. The fibrinogenolytic activity in human plasma was measured for wild-type, delta FE, delta FE1X, and delta FE3X t-PA, and showed significant fibrinogen depletion after 3 h of incubation at 51 nM, decreasing to 11, 11, 50, and 72% of basal levels, respectively. These data indicate that partial or total nonglycosylated t-PA variants have a higher fibrinolytic versus fibrinogenolytic ratio than their fully glycosylated counterparts.  相似文献   

6.
Recombinant human tissue plasminogen activator (rt-PA), produced by expression in Chinese hamster ovary cells, is a fibrin-specific plasminogen activator which has been approved for clinical use in the treatment of myocardial infarction. In this study, the structures of the Asn-linked oligosaccharides of Chinese hamster ovary-expressed rt-PA have been elucidated. High mannose and hybrid oligosaccharides were released from the protein by endoglycosidase H digestion, whereas N-acetyllactosamine-type ("complex") oligosaccharides were released by peptide:N-glycosidase F digestion. The oligosaccharides were fractionated by gel permeation chromatography and anion exchange high performance liquid chromatography (HPLC), and their structures were analyzed by composition and methylation analysis, high pH anion exchange chromatography, fast atom bombardment-mass spectrometry (FAB-MS), and 500-MHz 1H NMR spectroscopy. High mannose oligosaccharides were found to account for 38% of the total carbohydrate content of rt-PA and consisted of Man5GlcNAc2, Man6GlcNAc2, and Man7GlcNAc2 in the ratio 1.8:1.7:1. Two hybrid oligosaccharides were identified and accounted for 3% of the carbohydrate of rt-PA. The N-acetyllactosamine-type oligosaccharides were found to comprise diantennary (34% of total carbohydrate), 2,4-branched triantennary (11%), 2,6-branched triantennary (9%), and tetraantennary (5%) structures. Sialylation of these oligosaccharides was by alpha (2----3) linkages to galactose. Most (greater than 90%) of the N-acetyllactosamine-type structures contained fucose alpha (1----6) linked to the Asn-linked N-acetylglucosamine residue. The distribution of oligosaccharide structures at individual glycosylation sites (Asn residues 117, 184, and 448) was also determined. rt-PA exists as two variants that differ by the presence (type I) or absence (type II) of carbohydrate at Asn-184. Tryptic glycopeptides were isolated by reversed phase high performance liquid chromatography and treated with peptide:N-glycosidase F. The oligosaccharides released from each glycosylation site were analyzed by high pH anion exchange chromatography. By this analysis, Asn-117 was demonstrated to carry exclusively high mannose oligosaccharides. When glycosylated, Asn-184 carried diantennary, 2,4-branched triantennary, 2,6-branched triantennary, and tetraantennary N- acetyllactosamine oligosaccharides in the ratio 9.0:4.5:1.4:1. Asn- 448 carried the same types of oligosaccharides, but in the ratio 7.5:1.6:2.1:1. The distributions of Asn-linked oligosaccharides at positions 117 and 448 were found not to be affected by the presence or absence of carbohydrate at position 184. The relevance of the  相似文献   

7.
Electrophoretic analysis of endoglycosidase-treated tissue plasminogen activator obtained from human melanoma cells showed that the heterogeneity observed for the protein in these preparations is caused by an N-glycosidically linked N-acetyllactosamine type of carbohydrate chain which is present in about 50% of the molecules. An oligomannose type and an N-acetyllactosamine type of glycan is present in all molecules. Three glycopeptides were isolated and characterized by 1H-NMR, sugar determination, methylation analysis and amino acid determination. The exact attachment site for each of the three glycans could be deduced from the amino acid compositions of the glycopeptides. Asn-117 carries the oligomannose type of glycan, the structure of which was completely determined. Asn-184 is the site where the presence or absence of a biantennary N-acetyllactosamine type of glycan causes the size heterogeneity. The third N-glycosylation site, Asn-448, was found to carry a triantennary or tetraantennary N-acetyllactosamine type of carbohydrate chain.  相似文献   

8.
The enzyme tissue-type plasminogen activator (t-PA) and its substrate Glu-plasminogen can both bind to fibrin. The assembly of these three components results in about a 1000-fold acceleration of the conversion of Glu-plasminogen into plasmin. Fibrin binding of t-PA is mediated both by its finger (F) domain and its kringle-2 domain. Fibrin binding of Glu-plasminogen involves its kringle structures (K1-K5). It has been suggested that particular kringles contain lysine-binding sites and/or aminohexyl-binding sites, exhibiting affinity for specific carboxyl-terminal lysines and intrachain lysines, respectively. We investigated the possibility that t-PA and Glu-plasminogen kringles share common binding sites in fibrin, limitedly digested with plasmin. For that purpose we performed competition experiments, using conditions that exclude plasmin formation, with Glu-plasminogen and either t-PA or two deletion mutants, lacking the F domain (t-PA del.F) or lacking the K2 domain (t-PA del.K2). Our data show that fibrin binding of t-PA, mediated by the F domain, is independent of Glu-plasminogen binding. In contrast, partial inhibition by Glu-plasminogen of t-PA K2 domain-mediated fibrin binding is observed that is dependent on carboxyl-terminal lysines, exposed in fibrin upon limited plasmin digestion. Half-maximal competition of fibrin binding of both t-PA and t-PA del.F is obtained at 3.3 microM Glu-plasminogen. The difference between this value and the apparent dissociation constant of Glu-plasminogen binding to limitedly digested fibrin (12.1 microM) under these conditions is attributed to multiple, simultaneous interactions, each having a separate affinity. It is concluded that t-PA and Glu-plasminogen can bind to the same carboxyl-terminal lysines in limitedly digested fibrin, whereas binding sites composed of intrachain lysines are unique both for the K2 domain of t-PA and the Glu-plasminogen kringles.  相似文献   

9.
The effects of hypotensive agents (captopril, enalaprilat, and lisinopril) on the activities of components of the fibrinolytic system (FS) and the effects of antifibrinolytic agents (6-aminohexanoic acid (6-AHA) and tranexamic acid (t-AMCHA)) on the activities of angiotensin converting enzyme (ACE) were studied in vitro. Enalaprilat did not affect the FS activity. Captopril considerably inhibited the amidase activities of urokinase (u-PA), tissue plasminogen activator (t-PA), and plasmin ([I]50 (2.0?2.6) ± 0.1 mM), and the activation of Glu-plasminogen by t-PA and u-PA ([I]50 (1.50?1.80) ± 0.06 mM), which may be due to the presence of a mercapto group in the inhibitor molecule. Lisinopril did not affect the amidase activities of FS enzymes, but stimulated Glu-plasminogen activation by u-PA and inhibited activation fibrin-bound Glu-plasminogen by t-PA ([I]50 (12.0 ± 0.5) mM). Presumably, these effects can be explained by the presence in lisinopril of a Lys side residue, whose binding to lysine-binding Glu-plasminogen centers resulted, on the one hand, in the transformation from its closed conformation to a semi-open one and, on the other hand, in its desorption from fibrin. Unspecific inhibition of the activity of ACE, a key enzyme of the renin-angiotensin system, in the presence of 6-AHA and t-AMCHA ([I]50 10.0 ± 0.5 and 7.5 ± 0.4 mM, respectively) was found. A decrease in the ACE activity along with the growth of the fibrin monomer concentration was revealed. The data demonstrate that, along with endogenous mediated interaction between FS and RAS, relations based on the direct interactions of exogenous inhibitors of one system affecting the activities of components of another system can take place.  相似文献   

10.
Matrix metalloproteinase-3 (MMP-3 or stromelysin-1) specifically binds to tissue-type plasminogen activator (t-PA), without however, hydrolyzing the protein. Binding affinity to proMMP-3 is similar to single chain t-PA, two chain t-PA and active site mutagenized t-PA (Ka of 6.3 x 106 to 8.0 x 106 M-1), but is reduced for t-PA lacking the finger and growth factor domains (Ka of 2.0 x 106 M-1). Activation of native Glu-plasminogen by t-PA in the presence of proMMP-3 obeys Michaelis-Menten kinetics; at saturating concentrations of proMMP-3, the catalytic efficiency of two chain t-PA is enhanced 20-fold (kcat/Km of 7.9 x 10-3 vs. 4.1 x 10-4 microM-1.s-1). This is mainly the result of an enhanced affinity of t-PA for its substrate (Km of 1.6 microM vs. 89 microM in the absence of proMMP-3), whereas the kcat is less affected (kcat of 1.3 x 10-2 vs. 3.6 x 10-2 s-1). Activation of Lys-plasminogen by two chain t-PA is stimulated about 13-fold at a saturating concentration of proMMP-3, whereas that of miniplasminogen is virtually unaffected (1.4-fold). Plasminogen activation by single chain t-PA is stimulated about ninefold by proMMP-3, whereas that by the mutant lacking finger and growth factor domains is stimulated only threefold. Biospecific interaction analysis revealed binding of Lys-plasminogen to proMMP-3 with 18-fold higher affinity (Ka of 22 x 106 M-1) and of miniplasminogen with fivefold lower affinity (Ka of 0.26 x 106 M-1) as compared to Glu-plasminogen (Ka of 1.2 x 106 M-1). Plasminogen and t-PA appear to bind to different sites on proMMP-3. These data are compatible with a model in which both plasminogen and t-PA bind to proMMP-3, resulting in a cyclic ternary complex in which t-PA has an enhanced affinity for plasminogen, which may be in a Lys-plasminogen-like conformation. Maximal binding and stimulation require the N-terminal finger and growth factor domains of t-PA and the N-terminal kringle domains of plasminogen.  相似文献   

11.
The vampire bat salivary plasminogen activator (BatPA) is virtually inactive toward Glu-plasminogen in the absence of a fibrin-like cofactor, unlike human tissue-type plasminogen activator (tPA) (the kcat/Km values were 4 and 470 M-1 s-1, respectively). In the presence of fibrin II, tPA and BatPA activated Glu-plasminogen with comparable catalytic efficiencies (158,000 and 174,000 M-1 s-1, respectively). BatPA's cofactor requirement was partially satisfied by polymeric fibrin I (54,000 M-1 s-1), but monomeric fibrin I was virtually ineffective (970 M-1 s-1). By comparison, a variety of monomeric and polymeric fibrin-like species markedly enhanced tPA-mediated activation of Glu-plasminogen. Fragment X polymer was 2-fold better but 9-fold worse as cofactor for tPA and BatPA, respectively, relative to fibrin II. Fibrinogen, devoid of plasminogen, was a 10-fold better cofactor for tPA than fibrinogen rigorously depleted of plasminogen, Factor XIII, and fibronectin; the enhanced stimulatory effect of the less-purified fibrinogen was apparently due to the presence of Factor XIII. By contrast, the two fibrinogen preparations were equally poor cofactors of BatPA-mediated activation of Glu-plasminogen. BatPA possessed only 23 and 4% of the catalytic efficiencies of tPA and two-chain tPA, respectively, in hydrolyzing the chromogenic substrate Spectrozyme tPA. However in the presence of fibrin II, BatPA and tPA exhibited similar kcat/Km values for the hydrolysis of Spectrozyme tPA. Our data revealed that BatPA, unlike tPA, displayed a strict and fastidious requirement for polymeric fibrin I or II. Consequently, BatPA may preferentially promote plasmin generation during a narrow temporal window of fibrin formation and dissolution.  相似文献   

12.
To probe the effects of N-glycosylation on the fibrin-dependent plasminogenolytic activity of tissue-type plasminogen activator (t-PA), we have expressed a human recombinant t-PA (rt-PA) gene in Chinese hamster ovary (CHO) cells and in a murine C127 cell line. The resulting rt-PA glycoproteins were isolated and their associated N-linked oligosaccharide structures determined by using a combination of high-resolution Bio-Gel P-4 gel filtration chromatography, sequential exoglycosidase digestion, and methylation analysis. The results show that CHO rt-PA is N-glycosylated differently from murine C127 derived rt-PA. Further, both rt-PA's are N-glycosylated differently from t-PA derived from a human colon fibroblast and the Bowes melanoma cell line (Parekh et al., 1989), confirming that N-glycosylation of the human t-PA polypeptide is cell-type-specific. Both CHO and murine rt-PA were fractionated on lysine-Sepharose chromatography. The N-glycosylation of the major forms was analyzed and their fibrin-dependent plasminogenolytic activity determined by using an indirect amidolytic assay with Glu-plasminogen and a chromogenic plasmin substrate. The results suggest that the various forms of rt-PA differ from one another with respect to the kinetics of their fibrin-dependent activation of plasminogen. Together, these data support the notion (Wittwer et al., 1989) that N-glycosylation influences the fibrin-dependent catalytic activity of t-PA and that t-PA when expressed in different cell lines may consist of kinetically and structurally distinct glycoforms.  相似文献   

13.
Tissue plasminogen activator, separated into variants I and II (differing in Mr by 2000-3000), was reduced and [14C]carboxymethylated. Fragments from cleavages with enzymes and cyanogen bromide (CNBr) were separated by reverse-phase high-performance liquid chromatography and subjected to sequence degradations. All seven CNBr fragments were purified and found to be compatible with the cDNA-derived amino acid sequence [Pennica, D., Holmes, W. E., Kohr, W. J., Harkins, R. N., Vehar, G. A., Ward, C. A., Bennett, W. F., Ylverton, E., Seeburg, P. H., Heynecker, H. L., Goeddel, D. V., & Collen, D. (1983) Nature (London) 301, 214-221]. Chemical characterization of 93% of the 527 residues recovered in 50 peptides confirmed the indirectly deduced primary structure of the protein. The tryptic peptide patterns from the two variants were found to differ for one peptide (T15). Since carbohydrate was present in this peptide for variant I and since a marked difference in chromatographic behavior for T15 was observed in variant II, we conclude that carbohydrate differences in this peptide (i.e., Asn-184 in the numbering system of the cDNA-derived amino acid sequence) are the explanation for the size differences between variants I and II. Carbohydrate was also found at two other positions in the protein, corresponding to Asn-117 and Asn-448. However, a fourth potential glycosylation site, Asn-218, is apparently not utilized for carbohydrate attachment. The enzyme is inactivated by diisopropyl phosphorofluoridate, which covalently modifies the serine residue corresponding to position 478, identifying this as the active site serine residue.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
S Stack  R D Gray  S V Pizzo 《Biochemistry》1991,30(8):2073-2077
Laminin is a large multidomain glycoprotein with diverse biological activities which include stimulation of neurite outgrowth, enhancement of tumor metastasis, and promotion of cell growth, adhesion, and differentiation. A 19 amino acid synthetic peptide derived from the E8 fragment of the laminin A chain (Cys-Ser-Arg-Ala-Arg-Lys-Gln-Ala-Ala-Ser-Ile-Lys-Val-Ala-Val-Ser-Ala-Asp -Arg- NH2) was identified which promotes metastasis and stimulates collagenase IV activity in the culture medium of B16 melanoma cells (Kanemoto et al., 1990). We report that this peptide, here designated LamA2091-2108, is also a potent stimulator of tissue plasminogen activator (t-PA)-catalyzed plasminogen activation, resulting in a 22-fold increase in the kcat/Km of the activation reaction. The activity of purified type I and type IV collagenase was inhibited by LamA2091-2108 with IC50 values of 3 and 43 microM, respectively. These data support an alternative mechanism for the appearance of collagenase activity in the culture media of melanoma cells, namely, that the peptide stimulates plasminogen activation, subsequently generating collagenase activity.  相似文献   

15.
Activation of human Glu-plasminogen, Lys-plasminogen and low-Mr plasminogen (lacking lysine-binding sites) by pro-urokinase (pro-UK), obtained from a human lung adenocarcinoma cell line (Calu-3, ATCC), obeys Michaelis-Menten kinetics. Activation occurs with a comparable affinity (Km 0.40-0.77 microM), while the catalytic rate constant (kcat) is comparable for Glu-plasminogen (0.0022s-1) and low-Mr plasminogen (0.0034 s-1), but is somewhat higher for Lys-plasminogen (0.0106 s-1). The rate of activation of plasminogen by pro-UK is not significantly influenced by the presence of 6-aminohexanoic acid, purified fragments LBS I or LBS II or histidine-rich glycoprotein, indicating that the high affinity of pro-UK for plasminogen is not mediated via the high-affinity lysine-binding site of plasminogen located in kringles 1-3 (LBS I) nor via the low-affinity lysine-binding site comprised within kringle 4 (LBS II). The site(s) in plasminogen involved in the high-affinity interaction with pro-UK thus appear to be located within the low-Mr plasminogen moiety.  相似文献   

16.
The Chl-protein complexes of three maize (Zea mays L.) mutants and one barley (Hordeum vulgare L.) mutant were analyzed using low temperature Chl fluorescence emissions spectroscopy and LDS-polyacrylamide gel electrophoresis. The maize mutants hcf-3, hcf-19, and hcf-114 all exhibited a high Chl fluorescence (hcf) phenotype indicating a disruption of the energy transfer within the photosynthetic apparatus. The mutations in each of these maize mutants affects Photosystem II. The barley mutant analyzed was the well characterized Chl b-less mutant chlorina-f2, which did not exhibit the hcf phenotype. Chlorina-f2 was used because no complete Chl b-less mutant of maize is available. Analysis of hcf-3, hcf-19, and hcf-114 revealed that in the absence of CP43, LHC II can still transfer excitation energy to CP47. These results suggest that in mutant membranes LHC II can interact with CP47 as well as CP43. This functional interaction of LHC II with CP47 may only occur in the absence of CP43, however, it is possible that LHC II is positioned in the thylakoid membranes in a manner which allows association with both CP43 and CP47.Abbreviations hcf high chlorophyll fluorescence - LDS lithium dodecyl sulfate - LHC II light-harvesting complex of Photosystem II - LHC I light-harvesting complex of Photosystem I - CPIa chlorophyll-protein complex consisting of LHC I and the PS I core complex - CPI chlorophyll-protein complex consisting of the PS I core complex - CP47 47 kDa chlorophyll-protein of the Photosystem II core - CP43 43 kDa chlorophyll-protein of the Photosystem II core - CP29 29 kDa chlorophyll-protein of Photosystem II - CP26 26 kDa chlorophyll-protein of Photosystem II - CP24 24 kDa chlorophyll-protein of Photosystem II - fp free pigments  相似文献   

17.
The mechanism of activation of human Glu-plasminogen by fibrin-bound tissue-type plasminogen activator (t-PA) in a plasma environment or in a reconstituted system was characterized. A heterogeneous system was used, allowing the setting of experimental conditions as close as possible to the physiological fibrin/plasma interphase, and permitting the separate analysis of the products present in each of the phases as a function of time. The generation of plasmin was monitored both by spectrophotometric analysis and by radioisotopic analysis with a plasmin-selective chromogenic substrate and radiolabelled Glu-plasminogen respectively. Plasmin(ogen)-derived products were identified by SDS/PAGE followed by autoradiography and/or immunoblotting. When the activation was performed in a plasma environment, the products identified on the fibrin surface were Glu-plasmin (90%) and Glu-plasminogen (10%), whereas in the soluble phase only complexes between Glu-plasmin and its fast-acting inhibitor were detected. Identical results were obtained with a reconstituted system comprising solid-phase fibrin, t-PA, Glu-plasminogen and and alpha 2-antiplasmin. In contrast, when alpha 2-antiplasmin was omitted from the solution, Lys-plasmin was progressively generated on to the fibrin surface (30%) and released to the soluble phase. In the presence of alpha 2-antiplasmin or in plasma, the amount of active plasmin generated on the fibrin surface was lower than in the absence of the inhibitor: in a representative experiment the initial velocity of plasmin generation was 2.8 x 10(-3), 2.0 x 10(-3) and 1.8 x 10(-3) (delta A405/min) for 200 nM-plasminogen, 200 nM-plasminogen plus 100 nM-alpha 2-antiplasmin and native plasma respectively. Our results indicate that in plasma or in a reconstituted purified system containing plasminogen and alpha 2-antiplasmin at a ratio similar to that found in plasma (1) the activation pathway of native Glu-plasminogen proceeds directly to the formation of Glu-plasmin, (2) Lys-plasminogen is not an intermediate of the reaction and therefore (3) Lys-plasmin is not the final active product. However, in the absence of the inhibitor, Lys-plasmin and probably Lys-plasminogen, which is more readily activated to plasmin than is Glu-plasminogen, are generated as well.  相似文献   

18.
The dissolution of blood clots by plasmin is normally initiated in vivo by the activation of plasminogen to plasmin through the activity of tissue plasminogen activator (t-PA). The rate of plasminogen activation can be stimulated several orders of magnitude by the presence of fibrin-related proteins. Here we describe the kinetic analysis of both recombinant human t-PA (wild-type) and a t-PA variant produced by site-directed mutagenesis in which the original sequence from amino acids 296 to 299, KHRR, has been altered to AAAA. This tetra-alanine variant form of t-PA, K296A/H297A/R298A/R299A t-PA, we refer to as "KHRR" t-PA here. The plasminogen activating kinetics of wild-type t-PA (Activase alteplase) showed a catalytic efficiency which changed over 100-fold dependent on the stimulator in the assay. The lowest rate was in the absence of a stimulator. The following stimulators showed increasing ability to accelerate the catalytic efficiency of the reaction: fibrinogen, fragments of fibrinogen obtained by digestion with plasmin, fibrin, and slightly degraded fibrin. This increase in efficiency was driven primarily by decreases in the Michaelis constant (KM) of the reaction, whereas the catalytic rate constant (kcat) of the reaction did not change significantly. The "KHRR" variant of t-PA displayed novel kinetics with all stimulators tested. In the absence of a stimulator or with the poorer stimulators (fibrinogen and fibrinogen fragments), the KM values of the reaction with Activase alteplase and "KHRR" t-PA were similar. The kcat however, was lower with "KHRR" t-PA than with wild-type t-PA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
We constructed two human tissue-type plasminogen activator/urokinase (t-PA/u-PA) hybrid cDNAs which were expressed by transfection of mouse Ltk- cells. The properties of the secreted proteins were compared with those of recombinant t-PA (rt-PA) and high molecular weight (HMW) u-PA. The hybrid proteins each contain the amino-terminal fibrin-binding chain of t-PA fused to the carboxy-terminal serine protease moiety of u-PA but differ by a stretch of 13 amino acid residues between kringle 2 of t-PA and the plasmin cleavage site of u-PA. Hybrid protein rt-PA/u-PA I contains amino acids 1-262 of t-PA connected with amino acids 147-411 of u-PA, whereas hybrid protein rt-PA/u-PA II consists of the same t-PA segment and residues 134-411 of u-PA. We demonstrated fibrin binding for rt-PA, whereas the hybrid proteins bind to a lesser extent and HMW u-PA has no affinity for fibrin. Plasminogen activation by either one of the hybrid proteins in the absence of a fibrin substitute was similar to that by HMW u-PA, while rt-PA was much less active. The catalytic efficiency, in the presence of a fibrin substitute, increases more than 2000-fold for rt-PA, about 250-fold for hybrid proteins I and II, and 12-fold for HMW u-PA, respectively. Under these conditions the hybrid proteins are more efficient plasminogen activators than the parental ones. The hybrid molecules form a 1:1 molar complex with the human endothelial plasminogen activator inhibitor (PAI-1), analogous to that formed by rt-PA and HMW u-PA. The relative affinity of rt-PA for PAI-1 is 4.6-fold higher than that of HMW u-PA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Plasminogen, the zymogen form of the fibrinolytic enzyme plasmin, is known to undergo plasmin-mediated modification in vitro. The modified form, Lys-plasminogen, is superior to the native Glu-plasminogen in fibrin binding and as a substrate for activation by tissue-type plasminogen activator (t-PA). The present study was undertaken to determine the existence and significance of the Glu- to Lys-plasminogen conversion during t-PA-mediated lysis of plasma clots in vitro. When human plasma was supplemented with exogenous Lys-plasminogen and clotted, a dose-dependent shortening of lysis time was observed. Formation of Lys-plasminogen in situ during fibrinolysis was determined using 131I-Glu-plasminogen-supplemented plasma. By the time of lysis, Lys-plasminogen had accumulated to about 20% of the initial concentration of Glu-plasminogen. Quantitation of activation of both Glu- and Lys-plasminogen as well as the conversion of Glu- to Lys-plasminogen in plasma supplemented with both 131I-Glu-plasminogen and 125I-Lys-plasminogen was accomplished by determining the flux of the isotopically labeled species along three pathways: Glu-plasminogen-->Glu-plasmin, Glu-plasminogen-->Lys-plasminogen, and Lys-plasminogen-->Lys-plasmin. After a brief lag, the Glu-plasminogen activation rate was constant until lysis was achieved, at which point activation ceased. The Lys-plasminogen activation rate also was essentially constant until lysis but was not characterized by a lag phase. The rate of conversion of Glu- to Lys-plasminogen was nonlinear and correlated directly with the rate of fibrinolysis. By the time lysis had occurred, Glu-plasminogen consumption had been distributed equally between direct activation to plasmin and conversion to Lys-plasminogen, and 45% of the plasmin which had been formed was derived from Lys-plasminogen. These results demonstrate both the formation and the subsequent activation of Lys-plasminogen during fibrinolysis. As a result of improved fibrin binding and activation of Lys-plasminogen compared to Glu-plasminogen, the formation of Lys-plasminogen within a clot constitutes a positive feedback mechanism that can further stimulate the activation of plasminogen by t-PA as fibrinolysis progresses.  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