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.
Plasminogen activator inhibitor type-1 (PAI-1) is a major inhibitor of fibrinolysis by virtue of its capacity to inhibit urokinase-type plasminogen activator (uPA) and tissue-type plasminogen activator (tPA). Systemic inflammation is invariably associated with elevated circulating levels of PAI-1, and during human sepsis plasma PAI-1 concentrations predict an unfavorable outcome. Knowledge about the functional role of PAI-1 in a systemic inflammatory response syndrome is highly limited. In this study, we determined the role of endogenous PAI-1 in cytokine release induced by administration of LPS or staphylococcal enterotoxin B (SEB). Both LPS and SEB elicited secretion of PAI-1 into the circulation of normal wild-type (Wt) mice. Relative to Wt mice, PAI-1 gene-deficient (PAI-1(-/-)) mice demonstrated strongly elevated plasma IFN-gamma concentrations after injection of either LPS or SEB. In addition, PAI-1(-/-) splenocytes released more IFN-gamma after incubation with LPS or SEB than Wt splenocytes. Both PAI-1(-/-) CD4+ and CD8+ T cells produced more IFN-gamma upon stimulation with SEB. LPS-induced IFN-gamma release in mice deficient for uPA, the uPA receptor, or tPA was not different from IFN-gamma release in LPS-treated Wt mice. These results identify a novel function of PAI-1 during systemic inflammation, where endogenous PAI-1 serves to inhibit IFN-gamma release by a mechanism that does not depend on its interaction with uPA/uPA receptor or tPA.  相似文献   

2.
Extracellular proteolysis is considered to be required during neuritic outgrowth to control the adhesiveness between the growing neurite membrane and extracellular matrix proteins. In this work, PC12 nerve cells were used to study the modulation of proteolytic activity during neuronal differentiation. PC12 cells were found to contain and release a 70-75-kDa tissue-type plasminogen activator (tPA) and a much less abundant 48-kDa urokinase-type plasminogen activator. A plasminogen activator inhibitor (PAI) activity with molecular sizes of 54 and 58 kDa was also detected in PC12 cell conditioned medium and formed high-molecular-mass complexes with released tPA. Release of PAI activity was dependent on treatment with nerve growth factor (NGF), whereas tPA synthesis and release were under control of a cyclic AMP-dependent mechanism and increased on treatment with dibutyryl-cyclic AMP [(But)2cAMP] or cholera toxin. Simultaneous treatment with NGF and (But)2cAMP resulted in increases of both tPA and PAI release and enhancement of tPA-PAI complex formation. The resulting plasminogen activator activity in conditioned medium was high in (But)2cAMP-treated cultures with short neuritic outgrowth but remained low in NGF- or NGF plus (But)2cAMP-treated cultures, where neurite extension was, respectively, large and very large. These results suggest that excess proteolytic activity may be detrimental to neuritic outgrowth and that not only PAI release but also tPA-PAI complex formation is associated with production of large and stable neuritic outgrowth. This can be understood as an involvement of PAI in the protection against neurite-destabilizing proteolytic activity.  相似文献   

3.
mRNA levels for urokinase type plasminogen activator (uPA), tissue type plasminogen activator (tPA), plasminogen activator inhibitor-1 (PAI-1) and plasminogen activator inhibitor-2 (PAI-2) were examined in human diploid (neonatal foreskin) fibroblasts grown in 200-ml microcarrier suspension culture. Four different substrates were used. These included gelatin-coated polystyrene plastic, DEAE-dextran, glass-coated polystyrene plastic and uncoated polystyrene plastic. Our previous studies have shown that culture fluids from diploid fibroblasts grown on DEAE-dextran contained higher levels of plasminogen-dependent fibrinolytic activity than culture fluids from the same cells grown on other substrates. The increased plasminogen activator activity was due largely to elevated amounts of tPA (In Vitro Cell. Develop. Biol. 22: 575–582, 1986). The present study shows that there is a corresponding elevation of tPA mRNA in diploid fibroblasts cultured on DEAE-dextran relative to the other substrates. There does not appear to be any difference in uPA mRNA or in mRNA for PAI-1 or PAI-2 produced by the same cells on the four substrates. These data suggest that the influence of the substrate on plasminogen activator production is mediated at the genetic level.  相似文献   

4.
Low passage cultures of normal human keratinocytes produce several components of the plasminogen activator/plasmin proteolytic cascade, including urokinase plasminogen activator (uPA), tissue plasminogen activator (tPA), and two specific inhibitors. Studies here presented demonstrate that these cells also contain a high-affinity (Kd = 3 x 10(-10) M) plasma membrane-binding site for uPA. High molecular weight uPA, either as the single-chain precursor or two-chain activated form, bound to the receptor; however, low molecular weight (33 kD) uPA, tPA, or epidermal growth factor did not compete for binding, demonstrating specificity. Acid treatment, which removed endogenous uPA from the receptor, was required to detect maximal binding (45,000 sites per cell). To investigate the possibility that the uPA receptor on keratinocytes may be involved in epithelial migration during wound repair, cultures were wounded and allowed to migrate into the wounded site. Binding sites for uPA were localized by autoradiographic analysis of 125I-uPA binding as well as by immunocytochemical studies using anti-uPA IgG. With both techniques uPA binding sites were detected selectively on the plasma membrane of cells at the leading edge of the migrating epithelial sheet. This localization pattern suggests that uPA receptor expression on keratinocytes may be coupled to cell migration during cutaneous wounding.  相似文献   

5.
Human Tera 2 embryonal carcinoma cells switch gradually from rapidly growing undifferentiated cells to almost nonproliferating cells during retinoic acid (RA)-induced neuronal differentiation. This process is associated with the increased expression of type 1 plasminogen activator inhibitor (PAI 1) mRNA, and the secreted inhibitor is immobilized to the pericellular area. Furthermore, the differentiation is accompanied by a decrease in the amount of both the secreted tissue-type PA (tPA) and the mainly cell-associated urokinase-type PA (uPA) activity. In RA-differentiated cells, uPA becomes localized at the vinculin-rich cell-substratum adhesion sites. Fibroblast growth factor activity has been associated with various events during embryonal growth and with the regulation of proteolytic enzymes. A short-term treatment of the undifferentiated Tera 2 cells with basic fibroblast growth factor (bFGF) increases uPA mRNA levels and the cell-associated uPA activity, whereas the secretory tPA activity decreases. bFGF induces PAI 1 mRNA expression in the undifferentiated cells, but unlike PAI 1 protein after RA-treatment, the inhibitor does not accumulate around the cells but is released in the medium. A similar exposure to bFGF has less effect on the RA-differentiated Tera 2 cells. Under these conditions bFGF treatment leads to an increase in the amounts of PAI 1 and uPA mRNAs, but no changes in the localization of these components can be seen. Differentiation of human embryonal carcinoma cells is thus connected with an altered response to bFGF.  相似文献   

6.
We examined whether plasminogen activators (PAs) are produced by bovine cumulus-oocyte complexes (COCs) during maturation in vitro. The effects of epidermal growth factor (EGF) on production of PAs in oocytes and cumulus cells were also examined. When COCs were cultured for 24 h with 30 ng/ml EGF, three plasminogen-dependent lytic zones (58.5 +/- 3.5 kDa, 79.0 +/- 3.0 kDa, and 113.5 +/- 6.5 kDa) were observed. Addition of amiloride, a competitive inhibitor of urokinase-type PA (uPA), to the zymogram eliminated the activity of the 58.5 +/- 3.5-kDa zone, suggesting that this band is a uPA. However, since the activity of the remaining two bands was not eliminated, it was suggested that the 79.0 +/- 3.0-kDa band is a tissue-type PA (tPA) and the 113.5 +/- 6.5-kDa band is possibly a tPA-PA inhibitor (tPA-PAI) complex. In COCs before culture, however, no activity of PAs was detected. At 6 h of culture, the same level of uPA activity was detected in COCs cultured both in the absence and in the presence of EGF. The uPA activity was increased at 12 h of culture but without further increase at 24 h of culture, with higher activity in the presence than in the absence of EGF. The activity of tPA and tPA-PAI was first detected at 24 h of culture in the absence of EGF. In the presence of EGF, however, some activity of tPA-PAI was detected at 12 h of culture. At 24 h of culture, the activity of all PAs was detected in cumulus cells, but only uPA activity was detected in oocytes, with higher activity in the presence than in the absence of EGF. The uPA activity in oocytes was not detected when they were cultured without cumulus cells in either the presence or absence of EGF, although cumulus expansion was stimulated by EGF, exhibiting a time-course similar to that observed in PA production. These results suggest that uPA, tPA, and tPA-PAI are all produced by bovine COCs, but only uPA by oocytes, during maturation in vitro. However, cumulus cells play an essential role or roles in the production of uPA by oocytes, and EGF enhances the roles of cumulus cells.  相似文献   

7.
This study examined the effect of the preovulatory gonadotropin surge on the temporal and spatial regulation of tissue plasminogen activator (tPA), urokinase plasminogen activator (uPA), and uPA receptor (uPAR) mRNA expression and tPA, uPA, and plasmin activity in bovine preovulatory follicles and new corpora lutea collected at approximately 0, 6, 12, 18, 24, and 48 h after a GnRH-induced gonadotropin surge. Messenger RNAs for tPA, uPA, and uPAR were increased in a temporally specific fashion within 24 h of the gonadotropin surge. Localization of tPA mRNA was primarily to the granulosal layer, whereas both uPA and uPAR mRNAs were detected in both the granulosal and thecal layers and adjacent ovarian stroma. Activity for tPA was increased in follicular fluid and the preovulatory follicle apex and base within 12 h after the gonadotropin surge. The increase in tPA activity in the follicle base was transient, whereas the increased activity in the apex was maintained through the 24 h time point. Activity for uPA increased in the follicle apex and base within 12 h of the gonadotropin surge and remained elevated. Plasmin activity in follicular fluid also increased within 12 h after the preovulatory gonadotropin surge and was greatest at 24 h. Our results indicate that mRNA expression and enzyme activity for both tPA and uPA are increased in a temporally and spatially specific manner in bovine preovulatory follicles after exposure to a gonadotropin surge. Increased plasminogen activator and plasmin activity may be a contributing factor in the mechanisms of follicular rupture in cattle.  相似文献   

8.
Fibrinolysis is a process responsible for the dissolution of formed thrombi to re‐establish blood flow after thrombus formation. Plasminogen activator inhibitor‐1 (PAI‐1) inhibits urokinase‐type and tissue‐type plasminogen activator (uPA and tPA) and is the major negative regulator of fibrinolysis. Inhibition of PAI‐1 activity prevents thrombosis and accelerates fibrinolysis. However, a specific antagonist of PAI‐1 is currently unavailable for therapeutic use. We screened a panel of uPA variants with mutations at and near the active site to maximize their binding to PAI‐1 and identified a potent PAI‐1 antagonist, PAItrap. PAItrap is the serine protease domain of urokinase containing active‐site mutation (S195A) and four additional mutations (G37bR–R217L–C122A–N145Q). PAItrap inhibits human recombinant PAI‐1 with high potency (Kd = 0.15 nM) and high specificity. In vitro using human plasma, PAItrap showed significant thrombolytic activity by inhibiting endogenous PAI‐1. In addition, PAItrap inhibits both human and murine PAI‐1, allowing the evaluation in murine models. In vivo, using a laser‐induced thrombosis mouse model in which thrombus formation and fibrinolysis are monitored by intravital microscopy, PAItrap reduced fibrin generation and inhibited platelet accumulation following vascular injury. Therefore, this work demonstrates the feasibility to generate PAI‐1 inhibitors using inactivated urokinase.  相似文献   

9.
The urokinase plasminogen activator (uPA) and tissue plasminogen activator (tPA) are very similar serine proteases with the same physiological function, the activation of plasminogen. An increased amount or activity of uPA but not tPA has been detected in human cancers. The PAs are weak proteolytic enzymes, but they activate plasminogen to plasmin, a strong proteolytic enzyme largely responsible for the malignant properties of cancers. It has been shown recently that the administration of uPA inhibitors can reduce tumor size. Inhibitors of uPA could therefore be used as anti-cancer and anti-angiogenesis agents. It has been found that amiloride competitively inhibits the catalytic activity of uPA but not tPA. Modification of this chemical could therefore produce a new class of uPA specific inhibitors and a new class of anti-cancer agents. The X-ray structure of the uPA complex with amiloride is not known. There are structural differences in the specificity pocket of uPA and tPA. However, the potential energy of binding amiloride is lower outside this cavity in the case of tPA. A region responsible for binding amiloride to tPA has been proposed as the loop B93-B101, reached in negatively charged amino acids present in tPA but not uPA.  相似文献   

10.
The amidolytic plasmin activity of a mixture of tissue plasminogen activator (tPA) and plasminogen is enhanced by heparin at therapeutic concentrations. Heparin also increases the activity in mixtures of urokinase-type plasminogen activator (uPA) and plasminogen but has no effect on streptokinase or plasmin. Direct analyses of plasminogen activation by polyacrylamide gel electrophoresis demonstrate that heparin increases the activation of plasminogen by both tPA and uPA. Binding studies show that heparin binds to various components of the fibrinolytic system, with tight binding demonstrable with tPA, uPA, and Lys-plasminogen. The stimulation of tPA activity by fibrin, however, is diminished by heparin. The ability of heparin to promote plasmin generation is destroyed by incubation of the heparin with heparinase, whereas incubation with chondroitinase ABC or AC has no effect. Also, stimulation of plasmin formation is not observed with dextran sulfate or chondroitin sulfate A, B, or C. Analyses of heparin fractions after separation on columns of antithrombin III-Sepharose suggest that both the high-affinity and the low-affinity fractions, which have dramatically different anticoagulant activity, have similar activity toward the fibrinolytic components.  相似文献   

11.
It was shown previously that Ea4-peptide of trout pro-IGF-I exerted mitogenic activity in non-transformed cells and inhibited colony formation in a soft agar medium of established human cancer cells. Here we report that the same peptide inhibits the invasion of human breast cancer cells (MDA-MB-231) through a matrigel membrane in a dose-dependent manner. The expression of urokinase-type plasminogen activator (uPA), tissue-type plasminogen activator (tPA) and plasminogen activator inhibitor 1 (PAI1) genes in MDA-MB-231 cells were downregulated by treatment with rtEa4-peptide. The inhibition of expression of these genes in response to rtEa4-peptide treatment was reduced to the control level when inhibitors for c-Jun N-terminal kinase 1/2 (JNK1/2), mitogen activated protein kinase kinase 1/2 (Mek1/2), p38 mitogen activated protein kinase (p38 MAPK), phosphatidylinositol 3-kinase (PI3K), and phosphokinase C (PKC) were used. These results suggest that inhibition of invasion of MDA-MB-231 cells by rtEa4-peptide may be mediated via the suppression of uPA, tPA, and PAI1 gene activities through signal transduction pathways.  相似文献   

12.
This study evaluates the contribution of two types of plasminogen activators (PAs; tissue-type PA (tPA) versus urokinase-type PA (uPA) toward the invasiveness of human melanoma cells in a novel in vitro assay. We identified two human melanoma cell lines, MelJuso and MeWo, expressing uPA or tPA as shown at mRNA, protein, and enzyme activity level. MelJuso cells produced uPA as well as plasminogen activator inhibitor-1 (PAI-1). The latter was, however, not sufficient to neutralize the cell-associated or secreted uPA activity. MeWo cells secreted tPA, but the enzyme was not found to be cell-associated. PAI-1 production by these cells was not detectable. Plasminogen activation and fibrinolytic capacity of both cell lines were reduced by anticatalytic monoclonal antibodies specific for the respective type of PA or by aprotinin. In a novel in vitro invasion assay, antibodies to PA as well as aprotinin decreased the invasiveness of both cell lines into a fibrin gel, Matrigel, or intact extracellular matrix. Our results confirm the importance of uPA-catalyzed plasminogen activation in tumor cell invasiveness. Furthermore, we provide evidence that tPA, beyond its key role in thrombolysis, can also be involved in in vitro invasion of human melanoma cells.  相似文献   

13.
Summary A number of human endothelial cell lines from umbilical cord cells (HUVECs) have been generated by transfection with SV40 large T and small t antigen sequences. Comparison of these lines with primary cultures of HUVECs has been carried out by monitoring the expression of a number of endothelial cell markers with specific regard to cell age. The secreted levels of the protein plasminogen activator inhibitor (PAI) was found to be significantly reduced in SV40-transfected cells when compared to untransfected controls. Tissue plasminogen activator (tPA) and urokinase (uPA) levels were unchanged. As cells entered crisis, there was a rapid and significant increase in the levels of tPA, uPA, and PAI and this was observed for all clones screened. The endothelial cell marker von Willebrand Factor (vWF) was found intracellularly and was also secreted into the medium. The levels were not altered between transfected and untransfected cells. Angiotensin converting enzyme (ACE) activity was maintained in cell lines at levels found in nonimmortalized HUVECs. Both isoforms (α and β) of IL-1 (interleukin-1) increased as cells approached crisis, and the presence of these cytokines may be responsible for the increased levels of tPA, PAI, and uPA. With one exception, the ability of the transfected cells to produce prostacyclin (PGI2) was lost by all clones.  相似文献   

14.

Background

Proteolytic degradation by plasmin and metalloproteinases is essential for epidermal regeneration in skin wound healing. Plasminogen deficient mice have severely delayed wound closure as have mice simultaneously lacking the two plasminogen activators, urokinase-type plasminogen activator (uPA) and tissue-type plasminogen activator (tPA). In contrast, individual genetic deficiencies in either uPA or tPA lead to wound healing kinetics with no or only slightly delayed closure of skin wounds.

Methodology/Principal Findings

To evaluate the therapeutic potential in vivo of a murine neutralizing antibody directed against mouse uPA we investigated the efficacy in skin wound healing of tPA-deficient mice. Systemic administration of the anti-mouse uPA monoclonal antibody, mU1, to tPA-deficient mice caused a dose-dependent delay of skin wound closure almost similar to the delayed kinetics observed in uPA;tPA double-deficient mice. Analysis of wound extracts showed diminished levels of plasmin in the mU1-treated tPA-deficent mice. Immunohistochemistry revealed that fibrin accumulated in the wounds of such mU1-treated tPA-deficent mice and that keratinocyte tongues were aberrant. Together these abnormalities lead to compromised epidermal closure.

Conclusions/Significance

Our findings demonstrate that inhibition of uPA activity with a monoclonal antibody in adult tPA-deficient mice mimics the effect of simultaneous genetic ablation of uPA and tPA. Thus, application of the murine inhibitory mU1 antibody provides a new and highly versatile tool to interfere with uPA-activity in vivo in mouse models of disease.  相似文献   

15.
The reactions between plasminogen-activator inhibitor (PAI) and different plasminogen activators were studied in the presence of chromogenic peptide substrates for the enzymes. Our findings suggest that the rate constants for the reactions of PAI with single-chain tissue plasminogen activator (tPA), two-chain tPA, high-Mr urokinase and low-Mr urokinase are high and quite similar (1.6 X 10(7)-3.9 X 10(7) M-1.s-1). A free active site in the enzymes seems to be necessary for their reaction with PAI. Amino acids with antifibrinolytic properties did not interfere with the reactions. However, di-isopropyl phosphorofluoridate-inactivated tPA inhibited the reaction between PAI and all plasminogen activators in a similar way. These findings clearly demonstrated that a 'second-site' interaction, in addition to that between the enzyme active site and the inhibitor 'bait' peptide bond, is of importance for the high reaction rate. The reaction rate between PAI and single-chain tPA in the presence of an activator substrate (D-Ile-Pro-Arg p-nitroanilide) was decreased in the presence of fibrin. Fibrin caused a decrease in the Km for the single-chain tPA-substrate reaction. As a consequence, the 'free' concentration of single-chain tPA in the system decreased in the presence of fibrin, affecting the reaction rate between PAI and single-chain tPA. The phenomenon might be of physiological relevance, in the sense that single-chain tPA bound to fibrin in the presence of plasminogen would be protected against inactivation by PAI.  相似文献   

16.
Urokinase-type plasminogen activator (uPA) plays a central role in tissue remodeling processes. Most of our understanding of the role of uPA in vivo is derived from studies using gene-targeted uPA-deficient mice. To enable in vivo studies on the specific interference with uPA functionality in mouse models, we have now developed murine monoclonal antibodies (mAbs) directed against murine uPA by immunization of uPA-deficient mice with the recombinant protein. Guided by enzyme-linked immunosorbent assay, Western blotting, surface plasmon resonance, and enzyme kinetic analyses, we have selected two highly potent and inhibitory anti-uPA mAbs (mU1 and mU3). Both mAbs recognize epitopes located on the B-chain of uPA that encompasses the catalytic site. In enzyme activity assays in vitro, mU1 blocked uPA-catalyzed plasminogen activation as well as plasmin-mediated pro-uPA activation, whereas mU3 only was directed against the first of these reactions. We additionally provide evidence that mU1, but not mU3, successfully targets uPA-dependent processes in vivo. Hence, systemic administration of mU1 (i) rescued mice treated with a uPA-activable anthrax protoxin and (ii) impaired uPA-mediated hepatic fibrinolysis in tissue-type plasminogen activator (tPA)-deficient mice, resulting in a phenotype mimicking that of uPA;tPA double deficient mice. Importantly, this is the first report demonstrating specific antagonist-directed targeting of mouse uPA at the enzyme activity level in a normal physiological process in vivo.  相似文献   

17.
The influence of angiostatin K1-4.5--a fragment of the heavy chain of plasmin and a powerful inhibitor of angiogenesis--on kinetic parameters (k(Pg) and K(Pg)) of human Glu-plasminogen activation under the action of urokinase (uPA) not having affinity for fibrin and fibrin-specific tissue plasminogen activator (tPA) was investigated. Angiostatin does not affect the k(Pg) value, but increases the value K(Pg) urokinase plasminogen activation. A decrease in the k(Pg) value and an increase in the K(Pg) value were found for fibrin-stimulated plasminogen activation by tPA with increasing concentrations of angiostatin. The obtained results show that angiostatin is competitive inhibitor of the uPA activator activity, while it inhibits the activator activity of tPA by mixed type. Such an influence ofangiostatin on the kinetic constants ofthe urokinase plasminogen activation suggests that angiostatin dose dependent manner replaces plasminogen in the binary enzyme-substrate complex uPA-Pg. In case of fibrin-stimulated plasminogen activation by tPA, both zymogen and tPA are bound to fibrin with formation of the effective triple tPA-Pg-fibrin complex. Angiostatin replaces plasminogen both from the fibrin surface and from the enzyme-substrate tPA-Pg complex that leads to a decrease in k(Pg) and an increase in K(Pg) of plasminogen activation. Inhibition constants by angioststin (Ki) of plasminogen-activator activities of uPA and tPA determined by Dixon method were found to be 0.59 +/- 0.04 and 0.12 +/- 0.05 microM, respectively.  相似文献   

18.
19.
Plasminogen activator inhibitor-1 (PAI-1) is a specific inhibitor of the serine proteases tissue-type plasminogen activator (tPA) and urokinase-type plasminogen activator (uPA). To systematically investigate the roles of the reactive center P1 and P1' residues in PAI-1 function, saturation mutagenesis was utilized to construct a library of PAI-1 variants. Examination of 177 unique recombinant proteins indicated that a basic residue was required at P1 for significant inhibitory activity toward uPA, whereas all substitutions except proline were tolerated at P1'. P1Lys variants exhibited lower inhibition rate constants and greater sensitivity to P1' substitutions than P1Arg variants. Alterations at either P1 or P1' generally had a larger effect on the inhibition of tPA. A number of variants that were relatively specific for either uPA or tPA were identified. P1Lys-P1'Ala reacted 40-fold more rapidly with uPA than tPA, whereas P1Lys-P1'Trp showed a 6.5-fold preference for tPA. P1-P1' variants containing additional mutations near the reactive center demonstrated only minor changes in activity, suggesting that specific amino acids in this region do not contribute significantly to PAI-1 function. These findings have important implications for the role of reactive center residues in determining serine protease inhibitor (serpin) function and target specificity.  相似文献   

20.
Two types of plasminogen activator (tissue-type, tPA; urokinase-type, uPA) have been demonstrated in ovarian granulosa cells, but only tPA activity was found in denuded oocytes. Immature rats were treated subcutaneously with 20 IU pregnant mare's serum gonadotropin (PMSG) to stimulate follicle maturation, followed 2 days later by an injection of 10 IU human chorionic gonadotropin (hCG) to induce ovulation. Cellular plasminogen activator activities were determined by sodium dodecyl sulfate (SDS)-polyacrylamide gel electrophoresis followed by a fibrin-overlay technique. Cumulus-oocyte complexes from rats before and after PMSG treatment contained low amounts of tPA, but not uPA, activity. After hCG treatment, tPA activity showed a time-dependent increase, reaching a maximum at 24 h after injection. At 12 and 24 h after hCG treatment, uPA activity was also detected. The appearance of high molecular weight lysis zones further suggested the formation of plasminogen activator-inhibitor complexes. Morphological analysis indicated that the increases in oocyte tPA activity were correlated with the extent of cumulus cell expansion and dispersion. In denuded oocytes, tPA activity also progressively increased during the periovulatory period to a maximum at 24 h after hCG treatment. In contrast, neither uPA activity nor activator-inhibitor complex was detected. Secretion of the proteases was measured in the conditioned media of cumulus-oocyte complexes cultured for 24 h in vitro. Substantial increases in tPA release were found in complexes obtained at 8 and 12 h after hCG injection, with lower secretion from complexes obtained at 24 h after hCG treatment.(ABSTRACT TRUNCATED AT 250 WORDS)  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