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.
The glucose-regulated protein 78 (GRP78) is a plasminogen (Pg) receptor on the cell surface. In this study, we demonstrate that GRP78 also binds the tissue-type plasminogen activator (t-PA), which results in a decrease in Km and an increase in the Vmax for both its amidolytic activity and activation of its substrate, Pg. This results in accelerated Pg activation when GRP78, t-PA, and Pg are bound together. The increase in t-PA activity is the result of a mechanism involving a t-PA lysine-dependent binding site in the GRP78 amino acid sequence 98LIGRTWNDPSVQQDIKFL115. We found that GRP78 is expressed on the surface of neuroblastoma SK-N-SH cells where it is co-localized with the voltage-dependent anion channel (VDAC), which is also a t-PA-binding protein in these cells. We demonstrate that both Pg and t-PA serve as a bridge between GRP78 and VDAC bringing them together to facilitate Pg activation. t-PA induces SK-N-SH cell proliferation via binding to GRP78 on the cell surface. Furthermore, Pg binding to the COOH-terminal region of GRP78 stimulates cell proliferation via its microplasminogen domain. This study confirms previous findings from our laboratory showing that GRP78 acts as a growth factor-like receptor and that its association with t-PA, Pg, and VDAC on the cell surface may be part of a system controlling cell growth.  相似文献   

2.
Fibrin (Fn) clots formed from γ′-fibrinogen (γ′-Fg), a variant with an elongated γ-chain, are resistant to lysis when compared with clots formed from the predominant γA-Fg, a finding previously attributed to differences in clot structure due to delayed thrombin-mediated fibrinopeptide (FP) B release or impaired cross-linking by factor XIIIa. We investigated whether slower lysis of γ′-Fn reflects delayed plasminogen (Pg) binding and/or activation by tissue plasminogen activator (tPA), reduced plasmin-mediated proteolysis of γ′-Fn, and/or altered cross-linking. Clots formed from γ′-Fg lysed more slowly than those formed from γA-Fg when lysis was initiated with tPA/Pg when FPA and FPB were both released, but not when lysis was initiated with plasmin, or when only FPA was released. Pg bound to γ′-Fn with an association rate constant 22% lower than that to γA-Fn, and the lag time for initiation of Pg activation by tPA was longer with γ′-Fn than with γA-Fn. Once initiated, however, Pg activation kinetics were similar. Factor XIIIa had similar effects on clots formed from both Fg isoforms. Therefore, slower lysis of γ′-Fn clots reflects delayed FPB release, which results in delayed binding and activation of Pg. When clots were formed from Fg mixtures containing more than 20% γ′-Fg, the upper limit of the normal level, the delay in lysis was magnified. These data suggest that circulating levels of γ′-Fg modulate the susceptibility of clots to lysis by slowing Pg activation by tPA and provide another example of the intimate connections between coagulation and fibrinolysis.  相似文献   

3.
With the goal of identifying hitherto unknown surface exosites of streptokinase involved in substrate human plasminogen recognition and catalytic turnover, synthetic peptides encompassing the 170 loop (CQFTPLNPDDDFRPGLKDTKLLC) in the β-domain were tested for selective inhibition of substrate human plasminogen activation by the streptokinase-plasmin activator complex. Although a disulfide-constrained peptide exhibited strong inhibition, a linear peptide with the same sequence, or a disulfide-constrained variant with a single lysine to alanine mutation showed significantly reduced capabilities of inhibition. Alanine-scanning mutagenesis of the 170 loop of the β-domain of streptokinase was then performed to elucidate its importance in streptokinase-mediated plasminogen activation. Some of the 170 loop mutants showed a remarkable decline in kcat without any alteration in apparent substrate affinity (Km) as compared with wild-type streptokinase and identified the importance of Lys180 as well as Pro177 in the functioning of this loop. Remarkably, these mutants were able to generate amidolytic activity and non-proteolytic activation in “partner” plasminogen as wild-type streptokinase. Moreover, cofactor activities of the 170 loop mutants, pre-complexed with plasmin, against microplasminogen as the substrate showed a similar pattern of decline in kcat as that observed in the case of full-length plasminogen, with no concomitant change in Km. These results strongly suggest that the 170 loop of the β-domain of streptokinase is important for catalysis by the streptokinase-plasmin(ogen) activator complex, particularly in catalytic processing/turnover of substrate, although it does not seem to contribute significantly toward enzyme-substrate affinity per se.  相似文献   

4.
Angiostatin, a potent inhibitor of angiogenesis, is derived from the fibrinolytic proenzyme, plasminogen, by enzymatic processing. Plasminogen N-terminal activation peptide (PAP) is one of the products concomitantly released aside from angiostatin (kringles 1-4) and mini-plasminogen (kringle 5 plus the catalytic domain) when plasminogen is processed. To determine whether PAP alone or together with the angiostatin-related peptides derived from the processing of plasminogen modulate the proliferation and motility of endothelial cells, we have generated a recombinant PAP and used it to study its effects on endothelial cells in the presence and absence of the angiostatin-related peptides. Our results showed that PAP alone slightly increased the migration but not the proliferation of endothelial cells. However, in the presence of the angiostatin-related peptides, PAP attenuated the inhibitory activity of the angiostatin-related peptides on the proliferation and migration of endothelial cells. The inhibitory effect of PAP on the angiostatin-related peptides could be due to its binding to the kringle domains of the latter peptides.  相似文献   

5.
The Lyme disease spirochete Borrelia burgdorferi lacks endogenous, surface-exposed proteases. In order to efficiently disseminate throughout the host and penetrate tissue barriers, borreliae rely on recruitment of host proteases, such as plasmin(ogen). Here we report the identification of a novel plasminogen-binding protein, BBA70. Binding of plasminogen is dose-dependent and is affected by ionic strength. The BBA70-plasminogen interaction is mediated by lysine residues, primarily located in a putative C-terminal α-helix of BBA70. These lysine residues appear to interact with the lysine-binding sites in plasminogen kringle domain 4 because a deletion mutant of plasminogen lacking that domain was unable to bind to BBA70. Bound to BBA70, plasminogen activated by urokinase-type plasminogen activator was able to degrade both a synthetic chromogenic substrate and the natural substrate fibrinogen. Furthermore, BBA70-bound plasmin was able to degrade the central complement proteins C3b and C5 and inhibited the bacteriolytic effects of complement. Consistent with these functional activities, BBA70 is located on the borrelial outer surface. Additionally, serological evidence demonstrated that BBA70 is produced during mammalian infection. Taken together, recruitment and activation of plasminogen could play a beneficial role in dissemination of B. burgdorferi in the human host and may possibly aid the spirochete in escaping the defense mechanisms of innate immunity.  相似文献   

6.
We previously demonstrated that streptokinase (SK) can be used to generate active site-labeled fluorescent analogs of plasminogen (Pg) by virtue of its nonproteolytic activation of the zymogen. The method is versatile and allows stoichiometric and active site-specific incorporation of any one of many molecular probes. The limitation of the labeling approach is that it is both time-consuming and low yield. Here we demonstrate an improved method for the preparation of labeled Pg analogs by the use of an engineered SK mutant fusion protein with both COOH- and NH2-terminal His6 tags. The NH2-terminal tag is followed by a tobacco etch virus proteinase cleavage site to ensure that the SK Ile1 residue, essential for conformational activation of Pg, is preserved. The SK COOH-terminal Lys414 residue and residues Arg253–Leu260 in the SK β-domain were deleted to prevent cleavage by plasmin (Pm) and to disable Pg substrate binding to the SK·Pg/Pm catalytic complexes, respectively. Near elimination of Pm generation with the SKΔ(R253–L260)ΔK414–His6 mutant increased the yield of labeled Pg 2.6-fold and reduced the time required more than 2-fold. The versatility of the labeling method was extended to the application of Pg labeled with a near-infrared probe to quantitate Pg receptors on immune cells by flow cytometry.  相似文献   

7.
Plasma plasminogen is the precursor of the tumor angiogenesis inhibitor, angiostatin. Generation of angiostatin in blood involves activation of plasminogen to the serine protease plasmin and facilitated cleavage of two disulfide bonds and up to three peptide bonds in the kringle 5 domain of the protein. The mechanism of reduction of the two allosteric disulfides has been explored in this study. Using thiol-alkylating agents, mass spectrometry, and an assay for angiostatin formation, we show that the Cys462-Cys541 disulfide bond is already cleaved in a fraction of plasma plasminogen and that this reduced plasminogen is the precursor for angiostatin formation. From the crystal structure of plasminogen, we propose that plasmin ligands such as phosphoglycerate kinase induce a conformational change in reduced kringle 5 that leads to attack by the Cys541 thiolate anion on the Cys536 sulfur atom of the Cys512-Cys536 disulfide bond, resulting in reduction of the bond by thiol/disulfide exchange. Cleavage of the Cys512-Cys536 allosteric disulfide allows further conformational change and exposure of the peptide backbone to proteolysis and angiostatin release. The Cys462-Cys541 and Cys512-Cys536 disulfides have −/+RHHook and −LHHook configurations, respectively, which are two of the 20 different measures of the geometry of a disulfide bond. Analysis of the structures of the known allosteric disulfide bonds identified six other bonds that have these configurations, and they share some functional similarities with the plasminogen disulfides. This suggests that the −/+RHHook and −LHHook disulfides, along with the −RHStaple bond, are potential allosteric configurations.  相似文献   

8.
Cytokeratin 8 (CK8) is an intermediate filament protein that penetrates to the external surfaces of breast cancer cells and is released from cells in the form of soluble heteropolymers. CK8 binds plasminogen and tissue-type plasminogen activator (t-PA) and accelerates plasminogen activation on cancer cell surfaces. The plasminogen-binding site is located at the C-terminus of CK8. In this study, we prepared GST-fusion proteins which contained either 174 amino acids from the C-terminus of CK8 (CK8f) or 134 amino acids from the C-terminus of CK18 (CK18f). A third GST-CK fusion protein was identical to CK8f except that the C-terminal lysine was mutated to glutamine (CK8fK483Q). CK8f bound plasminogen; the K D was 0.5 M. Binding was completely inhibited by ACA. CK8fK483Q also bound plasminogen, albeit with decreased affinity (K D 1.5 M). CK18f did not bind plasminogen at all. All three fusion proteins bound t-PA equivalently, providing the first evidence that CK18 may function as a t-PA receptor. t-PA and plasminogen cross-competed for binding to CK8f. Thus, t-PA and plasminogen cannot bind to the same CK8f monomer simultaneously. Nevertheless, CK8f still promoted plasminogen activation, probably reflecting the fact that CK8f was purified in dimeric or tetrameric form. These studies demonstrate that CK8 may promote plasminogen activation by t-PA only when present in an oligomerized state. CK18 may participate in the oligomer, together with CK8, based on its ability to bind t-PA.  相似文献   

9.
Rapid kinetics demonstrate a three-step pathway of streptokinase (SK) binding to plasminogen (Pg), the zymogen of plasmin (Pm). Formation of a fluorescently silent encounter complex is followed by two conformational tightening steps reported by fluorescence quenches. Forward reactions were defined by time courses of biphasic quenching during complex formation between SK or its COOH-terminal Lys414 deletion mutant (SKΔK414) and active site-labeled [Lys]Pg ([5-(acetamido)fluorescein]-d-Phe-Phe-Arg-[Lys]Pg ([5F]FFR-[Lys]Pg)) and by the SK dependences of the quench rates. Active site-blocked Pm rapidly displaced [5F]FFR-[Lys]Pg from the complex. The encounter and final SK·[5F]FFR-[Lys]Pg complexes were weakened similarly by SK Lys414 deletion and blocking of lysine-binding sites (LBSs) on Pg kringles with 6-aminohexanoic acid or benzamidine. Forward and reverse rates for both tightening steps were unaffected by 6-aminohexanoic acid, whereas benzamidine released constraints on the first conformational tightening. This indicated that binding of SK Lys414 to Pg kringle 4 plays a role in recognition of Pg by SK. The substantially lower affinity of the final SK·Pg complex compared with SK·Pm is characterized by a ∼25-fold weaker encounter complex and ∼40-fold faster off-rates for the second conformational step. The results suggest that effective Pg encounter requires SK Lys414 engagement and significant non-LBS interactions with the protease domain, whereas Pm binding additionally requires contributions of other lysines. This difference may be responsible for the lower affinity of the SK·Pg complex and the expression of a weaker “pro”-exosite for binding of a second Pg in the substrate mode compared with SK·Pm.  相似文献   

10.
Streptokinase (SK) conformationally activates the central zymogen of the fibrinolytic system, plasminogen (Pg). The SK·Pg* catalytic complex binds Pg as a specific substrate and cleaves it into plasmin (Pm), which binds SK to form the SK·Pm complex that propagates Pm generation. Catalytic complex formation is dependent on lysine-binding site (LBS) interactions between a Pg/Pm kringle and the SK COOH-terminal Lys414. Pg substrate recognition is also LBS-dependent, but the kringle and SK structural element(s) responsible have not been identified. SK mutants lacking Lys414 with Ala substitutions of charged residues in the SK β-domain 250-loop were evaluated in kinetic studies that resolved conformational and proteolytic Pg activation. Activation of [Lys]Pg and mini-Pg (containing only kringle 5 of Pg) by SK with Ala substitutions of Arg253, Lys256, and Lys257 showed decreases in the bimolecular rate constant for Pm generation, with nearly total inhibition for the SK Lys256/Lys257 double mutant. Binding of bovine Pg (BPg) to the SK·Pm complex containing fluorescently labeled Pm demonstrated LBS-dependent assembly of a SK·labeled Pm·BPg ternary complex, whereas BPg did not bind to the complex containing the SK Lys256/Lys257 mutant. BPg was activated by SK·Pm with a Km indistinguishable from the KD for BPg binding to form the ternary complex, whereas the SK Lys256/Lys257 mutant did not support BPg activation. We conclude that SK residues Arg253, Lys256, and Lys257 mediate Pg substrate recognition through kringle 5 of the [Lys]Pg and mini-Pg substrates. A molecular model of the SK·kringle 5 complex identifies the putative interactions involved in LBS-dependent Pg substrate recognition.Streptokinase (SK)6 activates the human fibrinolytic system by activating plasminogen (Pg) through a unique mechanism that is responsible for the use of SK as a thrombolytic drug and its role as a key pathogenicity factor in Group A streptococcal infection (1, 2). The crystal structure of SK bound to the catalytic domain of plasmin (μPm) shows that SK consists of three β-grasp, tightly folded domains, α, β, and γ, linked by flexible segments (3). In solution, SK is highly flexible and behaves hydrodynamically like three beads on a string (4). When bound to μPm, SK assumes a highly ordered structure resembling a three-sided crater surrounding the catalytic site that provides an exosite(s) for binding the catalytic domain of Pg as a substrate (3, 5). In the first step of the SK-mediated Pg activation pathway, SK binds the catalytic domain of the Pg zymogen in a rapid equilibrium process and inserts its NH2-terminal Ile1 residue into the NH2-terminal binding cleft of Pg, activating the catalytic site nonproteolytically (610). Although structural proof is lacking, SK Ile1 presumably forms a critical salt bridge with Asp740(194) (plasminogen numbering; chymotrypsinogen numbering is in parentheses) that initiates conformational activation of the substrate binding site and oxyanion hole required for proteolytic activity (6, 810). The activated SK·Pg* complex binds a second molecule of Pg as a specific substrate and cleaves it at Arg561(15)-Val562(16) to form the fibrin-degrading proteinase, plasmin (Pm) (1014). Proteolytic generation of Pm is propagated by formation of a high affinity SK·Pm complex that converts the remaining free Pg into Pm (5, 11).[Glu]Pg, the full-length form of Pg circulating in blood, consists of an NH2-terminal PAN (Pg/Apple/Nematode (15, 16)) module, followed by five kringle domains (K1–K5), and the trypsin-like serine proteinase catalytic domain (17). Formation of the SK·Pg* and SK·Pm catalytic complexes and Pg substrate binding are inhibited by the lysine analog, 6-aminohexanoic acid (6-AHA), which binds to lysine-binding sites (LBS) located primarily in kringles K1, K4, and K5 of Pg and Pm (10, 11, 1823). Cleavage of the Lys77-Lys78 peptide bond in [Glu]Pg by Pm releases the PAN module and generates the truncated form, [Lys]Pg. Formation of [Lys]Pg is accompanied by a conformational change of [Glu]Pg from a compact, closed α-conformation to a partially extended β-conformation with expression of higher affinity LBS for 6-AHA (24, 25). The fourth kringle module mediates a second conformational change, from the β-conformation to the extended γ-conformation (25).Binding of SK to [Glu]Pg is independent of LBS, with a dissociation constant of 100–150 nm, whereas formation of SK·[Lys]Pg is LBS-dependent with a 13–20-fold higher affinity that is reduced to that of [Glu]Pg by saturating concentrations of 6-AHA (10, 21). Activation of the catalytic domain in [Lys]Pm increases affinity for SK about 830-fold, which is reduced 11–20-fold by 6-AHA (5, 21). Interaction of the COOH-terminal Lys414 residue of SK with a Pg/Pm kringle domain is responsible for the LBS-dependent enhancement of the affinity of SK·[Lys]Pg* and SK·Pm catalytic complex formation (22). Recent rapid reaction kinetic studies of the SK·Pm binding pathway demonstrated that interaction of Lys414 with a Pm kringle enhances formation of an initial rapid equilibrium SK·Pm encounter complex, succeeded by two sequential, tightening conformational changes, to achieve an overall dissociation constant of ∼12 pm (26). The Pg/Pm kringle domain responsible for the enhancement of SK·Pg* and SK·Pm complex formation is not known. Productive interaction of Pg as a substrate of the SK·Pg*/Pm complexes is also greatly inhibited by saturating 6-AHA (11). Kinetic and equilibrium binding studies of SK-mediated Pm formation resolved the conformational activation process from the coupled proteolytic generation of Pm (10, 11). The kinetic approach demonstrated that Lys414 deletion reduced the affinity of formation of the SK·Pg* catalytic complex specifically, whereas the subsequent LBS-dependent proteolytic formation of Pm was unaffected, indicating that Pg substrate recognition is mediated by a structurally distinct region of SK and an unknown kringle (22).Previous structure-function studies have yielded diverse interpretations and conclusions regarding the structural basis of LBS-dependent Pg substrate recognition (23, 2734). Each of the three domains of SK has been implicated in this regard (29, 30, 35, 36), and binding of two Pg molecules to the residue 1–59 sequence of the α-domain has been reported (36). In particular, segments 16–36, 41–48, 48–59, and 88–97 of the SK α-domain have been concluded to play a role in Pg substrate recognition (32, 33, 37, 38). For several SK mutants, a complex mixture of functional effects on their binding to [Glu]Pg and its conformational and proteolytic activation has been reported (28, 31, 33). Some of these effects may result from the inherent flexibility of SK when bound to Pg or Pm (39), and others may be due to the use of kinetic approaches that do not clearly discriminate between conformational and proteolytic activation.Some observations implicate a protruding hairpin loop called the 250-loop (residues Ala251–Ile264) in the SK β-domain in Pg substrate recognition (27, 28, 31, 34). This loop is disordered in the structure of the SK·μPm complex but is ordered in the structure of the isolated β-domain (3, 40). Deletion of the 250-loop, Ala substitution of Lys256 and Lys257 at the apex of the loop, and substitution of multiple residues near and within the loop resulted in disparate effects on Km and kcat for [Glu]Pg activation (27, 28, 31). The conclusions of these studies were that Lys256 and Lys257 are involved in SK binding and conformational activation of [Glu]Pg in addition to proteolytic processing of Pg as a substrate. Some of these studies are problematic because the natural NH2-terminal Ile1 residue necessary for conformational activation is preceded either by an additional methionine (27, 31) or maltose-binding protein (28) in the recombinant SK species used.Because of the diverse conclusions regarding the functional properties of the 250-loop mutations and the possibility of other potential Pg substrate binding sites, the present studies were undertaken to resolve the function of residues in the 250-loop in LBS-dependent Pg substrate recognition by the SK·Pg* complex. The kringle domain of Pg involved in Pg substrate recognition has not been clearly identified but has been suggested to be K5 (27) on the basis that the isolated β-domain bound Pg (30) and K5 (29) in an LBS-dependent manner. Given the general specificity of Pg kringles for COOH-terminal Lys residues and zwitterionic ligands, such as 6-AHA, and the internal sequence of the 250-loop, it appeared possible that a pseudolysine motif on SK was involved. In the binding of a 30-residue peptide from plasminogen binding Group A streptococcal M-like protein (PAM), VEK-30, to K2 of Pg, Castellino and co-workers (41, 42) showed by crystallography and mutagenesis that residues with cationic (Arg and His) and anionic side chains (Glu) arranged spatially on a helix constituted a pseudolysine structure similar to 6-AHA that binds specifically to the LBS of K2. Additional evidence for pseudolysine structures in Pg binding comes from studies of α-enolase from Streptococcus pneumoniae, which has a 9-residue internal binding site for Pg containing essential basic (two Lys residues) and acidic (Asp and Glu residues) located on a surface loop (43, 44).To determine whether a similar SK structure is involved in [Lys]Pg substrate recognition, anionic and cationic residues in the 250-loop were substituted with Ala and characterized in kinetic studies using methods that resolve conformational and proteolytic activation. Studies with [Lys]Pg and mini-Pg, which contains only K5 and the catalytic domain, showed that Arg253, Lys256, and Lys257 facilitate LBS-dependent substrate recognition through interactions with K5. The absence of evidence for a pseudolysine structure in the 250-loop is compatible with the established atypical specificity of K5 for cationic ligands, such as benzamidine, Nα-acetyl-Lys-methyl ester, 6-aminohexane, and 5-aminopentane, in addition to zwitterionic ligands (19, 4547). The studies resolve for the first time the structural features of SK that mediate the LBS-dependent interactions that enhance affinity of SK·Pg* and SK·Pm catalytic complex formation and those that facilitate binding of Pg as a substrate of these complexes.  相似文献   

11.
To gain insights into the mechanisms for the tight and highly specific interaction of the kringle 2 domain of human plasminogen (K2Pg) with a 30-residue internal peptide (VEK-30) from a group A streptococcal M-like protein, the dynamic properties of free and bound K2Pg and VEK-30 were investigated using backbone amide 15N-NMR relaxation measurements. Dynamic parameters, namely the generalized order parameter, S2, the local correlation time, τe, and the conformational exchange contribution, Rex, were obtained for this complex by Lipari-Szabo model-free analysis. The results show that VEK-30 displays distinctly different dynamic behavior as a consequence of binding to K2Pg, manifest by decreased backbone flexibility, particularly at the binding region of the peptide. In contrast, the backbone dynamics parameters of K2Pg displayed similar patterns in the free and bound forms, but, nonetheless, showed interesting differences. Based on our previous structure-function studies of this interaction, we also made comparisons of the VEK-30/K2Pg dynamics results from different kringle modules complexed with small lysine analogs. The differences in dynamics observed for kringles with different ligands provide what we believe to be new insights into the interactions responsible for protein-ligand recognition and a better understanding of the differences in binding affinity and binding specificity of kringle domains with various ligands.  相似文献   

12.
A monoclonal antibody to human plasminogen, 10-F-1, was found to interact with the lysine-binding site (LBS) on the kringle 4 (K 4) region of the molecule. This observation has been employed to measure the binding of various antifibrinolytic amino acid analogs of ?-aminocaproic acid (?ACA) to its site on K 4 in appropriate elastolytic-derived fragments of human plasminogen and to other species of plasminogen to which antibody 10-F-1 cross-reacts. By analysis of the concentration dependence of ?ACA displacement of [125I]10-F-1 from human Glu1Pg, a KD for ?ACA of 7.1 ± 1.0 mm was calculated. Similar experiments with K 4-containing fragments of Glu1Pg, viz., Lys77Pg, K 4, Lys77H and Val354Pg, yielded KD values of 6.6 ± 1.0, 7.5 ± 1.0, 6.6 ± 1.0, and 12.0 ± 2.0 mm, respectively. When baboon, goat, monkey, rabbit, and sheep plasminogens were substituted for human plasminogen, the KD values calculated ranged from 2.1 to 7.1 mm. The KD values for several analogs of ?ACA, i.e., 4-aminobutyric acid, 5-aminopentanoic acid, 8-aminooctanoic acid, l-lysine, and trans-aminomethyl cyclohexane-1-carboxylic acid, were measured to the K 4 region of Lys77Pg. The values obtained were 11.3 ± 1.5, 9.0 ± 1.0, 71.0 ± 10, 38.0 ± 5.0, and 1.1 ± 0.4 mm, respectively. Additionally, the KD of trans-aminomethylcyclohexane-1-carboxylic acid towards the K 4 region of Glu1Pg, Lys77Pg, and isolated K 4 was found to be 2.4 ± 0.5, 1.1 ± 0.3, and 2.0 ± 0.6 mm, respectively. These studies show directly that the LBS on the K 4 domain of plasminogen represents one of its 4–5 weak binding sites and that this site can be specifically probed with the use of monoclonal antibody 10-F-1. Furthermore, it appears as though this site is conserved in several important proteolytic fragments of plasminogen, providing additional evidence that these fragments exist as independent domains in the native molecule. Finally, this weak LBS on the K 4 domain of human plasminogen is also present in other species of plasminogen.  相似文献   

13.
Human plasminogen kringle 5 (K5) is known to display its potent anti-angiogenesis effect through inducing endothelial cell (EC) apoptosis, and the voltage-dependent anion channel 1 (VDAC1) has been identified as a receptor of K5. However, the exact role and underlying mechanisms of VDAC1 in K5-induced EC apoptosis remain elusive. In the current study, we showed that K5 increased the protein level of VDAC1, which initiated the mitochondrial apoptosis pathway of ECs. Our findings also showed that K5 inhibited the ubiquitin-dependent degradation of VDAC1 by promoting the phosphorylation of VDAC1, possibly at Ser-12 and Thr-107. The phosphorylated VDAC1 was attenuated by the AKT agonist, glycogen synthase kinase (GSK) 3β inhibitor, and siRNA, suggesting that K5 increased VDAC1 phosphorylation via the AKT-GSK3β pathway. Furthermore, K5 promoted cell surface translocation of VDAC1, and binding between K5 and VDAC1 was observed on the plasma membrane. HKI protein blocked the impact of K5 on the AKT-GSK3β pathway by competitively inhibiting the interaction of K5 and cell surface VDAC1. Moreover, K5-induced EC apoptosis was suppressed by VDAC1 antibody. These data show for the first time that K5-induced EC apoptosis is mediated by the positive feedback loop of “VDAC1-AKT-GSK3β-VDAC1,” which may provide new perspectives on the mechanisms of K5-induced apoptosis.  相似文献   

14.
A method of ELISA for measuring the binding of different samples of immunoglobulin (IgG) and its fragments to human plasminogen (Pg) has been developed. Instead of plasminogen, the heavy chain of plasminogen (Pg-H) containing five ligand-binding kringle domains, immobilized on the surface of the plate, was used in this method as a detector. It was found that IgG treated with plasmin (IgGPm-t) binds to the immobilized Pg-H 2.84 times more strongly than intact IgG. Both IgG samples showed a weak nonspecific binding to the immobilized light chain of plasminogen (Pg-L). It was shown that 0.2 M L-lysine inhibits the binding of IgGPm-t and does not affect the nonspecific binding of intact IgG to the immobilized Pg-H, indicating the involvement of lysine-binding regions of Pg-H in binding to IgGPm-t. A preliminary treatment of IgG samples with carboxypeptidase В (CPB) inhibited the binding of IgGPm-t and did not affect the nonspecific binding of intact IgG to the immobilized Pg-H, which indicates a key role of the С-terminal lysine of IgGPm-t in the specific binding to the lysine-binding sites of Pg. The study of the effects of intact IgG and IgGPm-t on the rate of activation of Glu- and Lys-forms of Pg (Glu-Pg and Lys-Pg) by a tissue activator of Pg (tPA) and urokinase (uPA) in buffer showed that intact IgG completely inhibited the activation of Glu-Pg and Lys-Pg with both tPA and uPA. Presumably, the inhibitory effect of intact IgG is due to steric hindrances that it creates for protein–protein interactions of the activators with the zymogen. IgGPm-t accelerated the generation of plasmin from Pg. In this case, the stimulatory effect of IgGPm-t on the activation of Glu-Pg under the action of tPA was ~25% higher than on the activation of Lys-Pg, which is explained by more significant conformational changes in the Glu-Pg molecule compared with the Lys-Pg molecule after their binding to IgGPm-t. The results suggest that the specific cleavage of IgG by plasmin may be one of the ways by which the plasminogen/plasmin system is involved in various physiological and pathological processes.  相似文献   

15.
Serum-free conditioned media and cell extracts from cultured human umbilical vein endothelial cells were analyzed for plasminogen activator by SDS-polyacrylamide gel electrophoresis and enzymography on fibrin-indicator gels. Active bands of free and complexed tissue-type plasminogen activator (t-PA) or urokinase-type plasminogen activator (u-PA) were identified by the incorporation of specific antibodies against, respectively, t-PA or u-PA in the indicator gel. The endothelial cells predominantly released a high-molecular-weight t-PA (95000–135000). This t-PA form was converted to Mr-72000 t-PA by 1.5 M NH4OH/39 mM SDS. A component with high affinity for both t-PA and u-PA could be demonstrated in serum-free conditioned medium and endothelial cell extract. The complex between this component and Mr-72000 t-PA comigrated with high-molecular-weight t-PA. From the increase in Mr of t-PA or u-PA upon complex formation, the Mr of the endothelial cell component was estimated to be 50000–70000. The reaction between t-PA or u-PA and the plasminogen activator-binding component was blocked by 5 mM p-aminobenzamidine, while the complexes, once formed, could be cleaved by 1.5 M NH4OH/39 mM SDS. These observations indicated that the active center of plasminogen activator was involed in the complex formation. It was further noted that serum-free conditioned medium of endothelial cell extract inhibited plasminogen activator activity when assayed by the fibrin-plate method. Evidence is provided that the plasminogen activator-binding component was different from a number of the known plasma serine proteinase inhibitors, the placenta inhibitor and the fibroblast surface protein, proteinase-nexin. We conclude that cultured endothelial cells produce a rapid inhibitor of u-PA and t-PA as well as a t-PA-inhibitor complex.  相似文献   

16.
Leishmania mexicana is able to interact with the fibrinolytic system through its component plasminogen, the zymogenic form of the protease plasmin. In this study a new plasminogen binding protein of this parasite was identified: LACK, the Leishmania homolog of receptors for activated C-kinase. Plasminogen binds recombinant LACK with a Kd value of 1.6 ± 0.4 μM, and binding is lysine-dependent since it is inhibited by the lysine analog ε-aminocaproic acid. Inhibition studies with specific peptides and plasminogen binding activity of a mutated recombinant LACK have highlighted the internal motif 260VYDLESKAV268, similar to those found in several enolases, as involved in plasminogen binding. Recombinant LACK and secreted proteins, in medium conditioned by parasites, enhance plasminogen activation to plasmin by the tissue plasminogen activator (t-PA). In addition to its localization in the cytosol, in the microsomal fraction and as secreted protein in conditioned medium, LACK was also localized on the external surface of the membrane. The results presented here suggest that LACK might bind and enhance plasminogen activation in vivo promoting the formation of plasmin. Plasminogen binding of LACK represents a new function for this protein and might contribute to the invasiveness of the parasite.  相似文献   

17.
In the course of analysis of plasminogen in microglial conditioned medium (Mic-CM), novel low-molecular-weight (LMW) zymogen with a molecular mass of ~36 kDa was detected by casein-urokinase zymography. Because this form was produced when rat native plasminogen was incubated with Mic-CM, a specific protease in the Mic-CM was thought to be responsible for the production of LMW plasminogen. The production of LMW plasminogen was strongly inhibited by elastase inhibitors. Furthermore, elastase (pancreatic or leukocyte) was also found to produce LMW zymogen from native plasminogen. These results indicate that LMW plasminogen is produced through limited proteolysis by an elastase-like protease in Mic-CM. To determine the biochemical characteristics of LMW plasminogen, rat native plasminogen was cleaved by pancreatic elastase, and the fragments (LMW plasminogen and nonzymogen fragments) were purified by several kinds of column chromatography. Amino acid sequence analysis revealed that LMW plasminogen is a carboxy-terminal region that contains the fifth kringle domain and a protease active site, and the amino acid sequence is identical to that of LMW plasminogen produced by MicCM. On the other hand, the nonzymogen fragment was the amino-terminal region containing four kringle domains. The effects of native plasminogen and the fragments on neurite outgrowth of rat brain explant were examined. LMW plasminogen promoted neurite outgrowth as well as did native plasminogen, whereas nonzymogen fragments did not. These results suggest that LMW plasminogen, which is produced from native plasminogen by elastase, may be a physiologically active molecule that mediates the intercellular interaction between microglia and neurons.  相似文献   

18.
Structure and function of human tissue-type plasminogen activator (t-PA)   总被引:5,自引:0,他引:5  
Full-length tissue-type plasminogen activator (t-PA) cDNA served to construct deletion mutants within the N-terminal "heavy" (H)-chain of the t-PA molecule. The H-chain cDNA consists of an array of structural domains homologous to domains present on other plasma proteins ("finger," "epidermal growth factor," "kringles"). These structural domains have been located on an exon or a set of exons. The endpoints of the deletions nearly coincide with exon-intron junctions of the chromosomal t-PA gene. Recombinant t-PA deletion mutant proteins were obtained after transient expression in mouse Ltk- cells, transfected with SV40-pBR322-derived t-PA cDNA plasmids. It is demonstrated that the serine protease moiety of t-PA and its substrate specificity for plasminogen is entirely contained within the C-terminal "light" (L)-chain of the protein. The presence of cDNA, encoding the t-PA signal peptide preceding the remaining portion of t-PA, suffices to achieve secretion of (mutant) t-PA into the medium. The stimulatory effect of fibrin on the plasminogen activator activity of t-PA was shown to be mediated by the kringle K2 domain and, to a lesser extent, by the finger domain. The other domains on the H-chain, kringle K1, and the epidermal growth-factor-like domain, do not contribute to this property of t-PA. These findings correlate well with the fibrin-binding properties of the rt-PA deletion-mutant proteins, indicating that stimulation of the activity is based on aligning of the substrate plasminogen and its enzyme t-PA on the fibrin matrix. The primary target for endothelial plasminogen activator inhibitor (PAI) is located within the L-chain of t-PA. Deleting specific segments of t-PA H-chain cDNA and subsequent transient expression in mouse Ltk- cells of t-PA deletion-mutant proteins did not affect the formation of a stable complex between mutant t-PA and PAI.  相似文献   

19.
The biochemical essence of prion replication is the molecular multiplication of the disease-associated misfolded isoform of prion protein (PrP), termed PrPSc, in a nucleic acid-free manner. PrPSc is generated by the protein misfolding process facilitated by conformational conversion of the host-encoded cellular PrP to PrPSc. Evidence suggests that an auxiliary factor may play a role in PrPSc propagation. We and others previously discovered that plasminogen interacts with PrP, while its functional role for PrPSc propagation remained undetermined. In our recent in vitro PrP conversion study, we showed that plasminogen substantially stimulates PrPSc propagation in a concentration-dependent manner by accelerating the rate of PrPSc generation while depletion of plasminogen, destabilization of its structure and interference with the PrP-plasminogen interaction hinder PrPSc propagation. Further investigation in cell culture models confirmed an increase of PrPSc formation by plasminogen. Although molecular basis of the observed activity for plasminogen remain to be addressed, our results demonstrate that plasminogen is the first cellular protein auxiliary factor proven to stimulate PrPSc propagation.Key words: prion, PrPSc, protein misfolding, auxiliary factor, plasminogen, PMCA, cell culture modelPrions are unique infectious particles that replicate in the absence of nucleic acids1 and cause fatal neurologic disorders in mammals.2 In fact, the prion particle is composed of an alternatively folded form of the cellular prion protein (PrPC) encoded by the Prnp gene. The misfolded form of PrPC, referred to as PrPSc, shares the same primary structure with PrPC,3 but exhibits distinctively different biochemical and biophysical properties.4,5 Moreover, animals lacking expression of the Prnp gene are resistant to prion infection,6 suggesting that PrPC serves as a precursor for PrPSc.The essence of prion biosynthesis is based on the protein-only hypothesis that postulates self-perpetuating replication of the prion protein.1 Although the exact replication process remains elusive, the template-assisted conversion model proposes an idea that PrPSc serves as a template to convert α-helices of PrPC into the β-sheets of PrPSc during prion replication.7 According to many lines of evidence, there exists an unidentified auxiliary factor, designated protein X, which favorably interacts with PrPC to produce a thermodynamically stable intermediate conformation called PrP*.8 By introducing PrPSc to a host, dimers will readily form between PrP* and PrPSc. This interaction induces PrP* to take on the conformation of PrPSc and results in a complex consisting of the template and the newly formed PrPSc molecules. Once the dimer disassociates protein X and the two PrPSc molecules are released and allowed to continue replicating in an exponential fashion. Recent observations that infectious material can be generated in vitro using recombinant PrP have authenticated the protein-only hypothesis.9,10 However, the infectivity of these in vitro generated PrPSc products is lower than that of brain-derived PrPSc, leading to the possibility that insufficient levels of the unidentified auxiliary factor are the limiting factor for these in vitro assays.To date, non-mammalian chaperone proteins, sulfated glycans and certain polyanionic macromolecules, such as RNA, have shown to increase the level of PrPSc in several different in vitro assays.1114 However, defining these molecules as cellular auxiliary factors that promote the conversion of PrPC to PrPSc has been prevented for several reasons. First, overexpression of yeast heat shock protein Hsp 104 in transgenic mice does not modulate the incubation time of disease and PrPSc accumulation upon prion inoculation.15 Although mammalian Hsp 70 is upregulated in humans and animals with prion diseases,16,17 it participates in downregulation, but not upregulation, of PrPSc accumulation.18 Second, the effect of sulfated glycans on PrPSc formation has been inconsistent1113 and certain sulfated glycans inhibit PrPSc propagation in animals and cultured cells.1921 Lastly, although RNA is able to increase PrPSc propagation and induce the formation of PrPSc de novo from purified PrPC in the absence of a PrPSc seed,22 its specificity is in question. Thus, an auxiliary factor that positively assists PrPSc replication and is composed of a mammalian cellular protein remains to be identified.Our approach to identify an auxiliary factor relies on the idea that the auxiliary factor interacts with PrPC as proposed in the protein X hypothesis.7 Therefore, we considered PrP ligands as the best candidates for this unidentified factor. By screening a phage display cDNA expression library from ScN2a cells, we identified kringle domains of plasminogen that interact with recombinant PrP folded in an α-helical conformation (α-PrP).23 In vitro binding assays showed that interaction between plasminogen and α-PrP that represents the conformational state of PrPC was enhanced by the introduction of a dominant negative mutation and the presence of the basic N-terminal sequences in PrP.23 Interaction of PrPC and plasminogen was further confirmed by the ability of plasminogen and its kringle domains to readily interact with α-PrP.2429 However, despite greater interaction with α-PrP, plasminogen also interacted with PrP in β-sheet conformations.23 Prior to our study, it was shown that PrPSc was immunoprecipitated with beads linked to plasminogen, its first three kringle domains [K(1+2+3)], and more recently the repeating YRG motif found within the plasminogen kringle domains.3033 However, the ability of plasminogen to bind to PrPSc was dependent on conditions of the lipid rafts and plasminogen was actually associated with PrPC in the intact lipid rafts.34To determine the functional relevance of this interaction for PrPSc replication, we explored whether plasminogen enhances PrPSc propagation using cell culture models and the in vitro PrP conversion assay, termed protein misfolding cyclic amplification (PMCA).35 The addition of plasminogen in PMCA resulted in the generation of significantly more PrPSc in a concentration-dependent manner (Fig. 1A), suggesting that plasminogen stimulates the conversion of PrPC to PrPSc. Indeed, our kinetic studies showed that plasminogen accelerates the rate of PrPSc generation during the early stages of the in vitro PrP conversion reaction (reviewed in ref. 35). In contrast, the addition of plasminogen in PMCA lacking either PrPC or PrPSc failed to generate PrPSc (Fig. 1B and C). These results suggest that both PrPC and PrPSc are required for PrPSc replication stimulated by plasminogen and that plasminogen facilitates neither spontaneous PrP conversion nor PrPSc augmentation through aggregating pre-existing PrPSc. Incubation of plasminogen with pre-formed PrPSc followed by treatments with proteinase K failed to either increase or decrease the PrPSc level compared to controls (Fig. 1D and E), suggesting that plasminogen is not involved in stabilization of PrPSc, enhancement of PrPSc resistance to protease or promotion of PrPSc binding to the membrane for western blotting. The activity to stimulate PrP conversion was a specific property of plasminogen and was not shared with other proteins such as a known PrP ligand and proteins abundantly found in the serum or at the extracellular matrices where plasminogen is present (reviewed in ref. 35). In addition, we found that the ability of plasminogen to assist in PrPSc propagation is preserved in its kringle domains (reviewed in ref. 35). Furthermore, the activity associated with plasminogen under cell-free conditions was reproduced in cell culture models. Plasminogen and its kringle domains increased PrPSc propagation in cultured cells chronically infected with mouse-adapted scrapie or chronic wasting disease prions (Fig. 2A–D). This suggests that the activity of plasminogen in PrPSc replication has biological relevance.Open in a separate windowFigure 1The role of plasminogen in PrPSc propagation. The effect of plasminogen (Plg) was assessed by PMCA using normal brain material supplemented with or without 0.5 µM human Glu-Plg (A–D) or using Plg-deficient (Plg-/-) brain material (F and G). Pre- (−) and post- (+) PMCA samples were treated with proteinase K (PK) and analyzed by western blotting. Seeds for PMCA were diluted either as indicated or 1:900 (G) −8,100 (F). (A) Stimulation of PrPSc propagation by Plg. (B) Plg-supplemented PMCA in the absence of SBH seeds. (C) Plg-supplemented PMCA in the absence of NBH. (D) Comparison of PrPSc levels in Plg-supplemented PMCA samples (during) vs. PMCA samples only incubated with Plg prior to PK digestion (after). (E) Comparison of PrPSc levels of ScN2a cell lysate after incubation with or without Plg prior to PK digestion. (F) PMCA with brain material of Plg-/- mice and genetically unaltered littermate controls (C). (G) Restoration of PMCA using Plg-/- brain material with Plg-supplementation. NBH, normal brain homogenate; SBH, sick brain homogenate; PrPKOBH, brain homogenate of PrPC-deficient mice; CL, cell lysate. Reproduced with permission from The FASEB Journal, Mays and Ryou 2010.35Open in a separate windowFigure 2PrPSc propagation increased by plasminogen in prion-infected cells. (A) The levels of PrP in ScN2a cells incubated with 0–0.5 µM human Glu-plasminogen (Plg) for two days. (B) The levels of PrP in ScN2a cells incubated with 0, 0.1 and 1.0 µM Plg or the first three kringle domains of Plg [K(1+2+3)] for six days. (C) The levels of 3F4-tagged PrPC and nascent PrPSc formation in ScN2a cells transiently transfected (Tfx) with plasmids encoding the 3F4-tagged PrP gene (PrP-3F4) and with an empty vector (mock). The transfected cells were treated with 0 or 1 µM K (1+2+3) for three days. (D) The levels of PrP in Elk21+ cells incubated with 0 and 0.5 µM Plg for two days. PrP was detected by anti-PrP antibody D13 (A and B), 3F4 (C) or 6H4 (D) before (−) and after (+) PK treatment. Reproduced by permission of the The FASEB Journal, Mays and Ryou 2010.35Corresponding to the results that plasminogen positively assists in PrPSc replication by stimulating conversion of PrPC to PrPSc, depletion of plasminogen from the PMCA reaction by using brain material derived from plasminogen-deficient mice restricted PrPSc replication to the basal level (Fig. 1F). Supplementation with plasminogen for PMCA using plasminogen-deficient brain homogenate restored PrPSc propagation to levels equivalent to that of control PMCA in which only brain material of their non-genetically altered littermates was used (Fig. 1G). Furthermore, structural destabilization of plasminogen affected the activity of plasminogen that enhances PrPSc propagation. Because intact disulfide bonds are critical in maintaining structural integrity and the binding activity of plasminogen, we conducted PMCA supplemented with either structurally intact or modified plasminogen to investigate the functionality of plasminogen. The result showed that plasminogen pre-treated sequentially with chemical agents that disrupt disulfide bonds and modify free sulfhydryl groups failed to stimulate PrPSc propagation (reviewed in ref. 35). In addition, we showed that interference with the plasminogen-PrP interaction using L-lysine abolished the plasminogen-mediated stimulation of PrPSc propagation in PMCA (reviewed in ref. 35). Because previous observations in immunoprecipitation30 and ELISA binding assays23 described that L-lysine specifically inhibited formation of the PrP-plasminogen complex, the presence of L-lysine in PMCA is considered to saturate the lysine binding motifs of kringle domains, which competitively prevents the kringle domains of plasminogen from interacting with PrP and inhibited PrP conversion. These various inhibition studies of PrPSc propagation provides confirmatory evidence that plasminogen plays an important role in PrPSc replication.Despite our progress in understanding the role of plasminogen in PrPSc propagation, we are still unable to address mechanistic details by which plasminogen exerts its function. In fact, plasminogen shares a number of the expected characteristics of the previously proposed auxiliary factor, although there are minor but distinctive discrepancies in their properties (summarized in Fig. 3). First, plasminogen may control conformational rearrangement of PrPC to PrP*, resembling a molecular chaperone. This scenario is identical to the protein X hypothesis, in which plasminogen replaces protein X to assist the conversion process. Second, plasminogen may promote aggregation of PrPSc already converted from PrPC. This will result in efficient formation of PrPSc multimers. Third, plasminogen may stabilize the pre-existing PrPSc aggregates so that stabilized aggregates are better protected from degradation or processing by the intrinsic clearance mechanism. This will result in increased accumulation and stability of PrPSc. In the second and third scenarios, the function of plasminogen is not involved in PrPC or its conversion process, but limited to the interaction with pre-formed PrPSc. Fourth, plasminogen may play a role as a scaffolding molecule that simply brings both PrPC and PrPSc together within a proximity. This will increase the frequency of interaction between PrPC and PrPSc for conversion. This scenario is distinguished from the other three by postulating plasminogen interaction with both isoforms of PrP. In contrast, plasminogen is assumed to interact with only PrPC or PrPSc in other cases. Although accumulating data including our recent studies provide critical pieces of evidence to envision described mechanistic insights, it is still premature to conclude the mechanism involved in plasminogen-mediated stimulation of PrPSc propagation.Open in a separate windowFigure 3Plausible mechanisms for plasminogen to enhance PrPSc propagation. Plasminogen may stimulate PrPSc propagation via conformational alteration of PrPC to PrP* (i), enhancement of PrPSc aggregation (ii), stabilization of pre-exisiting PrPSc aggregates (iii) or scaffolding to gather PrPC and PrPSc together (iv).

Table 1

Properties of plasminogen as an auxiliary factor for PrPSc propagation
Protein XPlasminogen
CompositionProtein; macromoleculesProtein
ExpressionBrain; neuron-specificBrain; neuroblastoma cell line, expressed more in the non-CNS
Subcellular localizationPlasma membrane; lipid raftsExtracellular matrix; lipid rafts
Association with diseaseIncreased protein levels in the sera of human patient with CJD
InteractionOnly with PrPCPrPSc, α-PrP, β-PrP
Binding sites on PrP
  • Q167, Q171, V214 and Q218 in the β2-α2 loop (164–174) and C-terminus (215–223)
  • Dominant negative mutations on the protein X binding sites such as Q167R and Q218K inhibited PrPSc formation in the cultured cells and prion transmission in transgenic mice
  • K23, K24 and K27 in the N-terminus: deletion of the N-terminal lysine cluster reduced dominant negative inhibition of PrPSc formation
  • Binds to N-terminally truncated PrP (89–230)
  • Increased binding activity to full-length PrP (23–230)
  • Increased binding activity to PrP with Q218K dominant negative mutation
  • The second kringle domain of plasminogen binds to the β2-α2 loop in silico
  • Binds to both lysine clusters located to the N-terminus (23–27) and middle (100–109) of PrP
Species specificity
  • Homotypic interaction with PrPC
  • Mouse protein X has lower binding affinity to human PrPC in the studies with transgenic mice
  • Unknown
  • Human plasminogen binds to human, mouse, bovine and ovine PrP
  • Human and bovine plasminogen converts mouse PrPC to PrPSc in PMCA
FunctionAn auxiliary role in conversion of PrPC to PrPScEnhances PrPSc propagation facilitated by PrP conversion in PMCA
Action mechanismBinds to PrPC and alters PrPC into PrP* that interacts with PrPSc for conversionUnknown
Open in a separate windowCNS, central nervous system; CJD, Creutzfeldt-Jakob disease; α-PrP, PrP in an α-helical conformation; β-PrP, PrP in an β-sheet conformation; PMCA, protein misfolding cyclic amplification.It is essential to address a few additional issues of PrPSc propagation stimulated by plasminogen aside from the mechanistic details. First, it is necessary to confirm the authenticity of PrPSc generated in the aid of plasminogen as a bona fide infectious agent. In addition, it would be interesting to compare the structural signatures of PrPSc generated in our study and found the prion seeds. This study may provide a clue to establish a structure-infectivity relationship for PrPSc. Furthermore, the ability to repeat our PMCA results in a similar assay reconstituted with defined components would clarify whether plasminogen directly contributes to PrPSc propagation. Finally, the contribution of plasminogen to the species barrier by controlling the compatibility between the host PrPC and prion strains should be determined. Collectively, these studies will be useful to identify an intricate regulatory role for plasminogen during PrPSc replication and prion transmission.Plasminogen has shown that it stimulates PrPSc propagation in the cell-free assays and cultured cells, while its role in animal models has not been obviously clarified. In opposition to the anticipated outcomes that the course of prion disease in plasminogen-deficient animals would be delayed, intracerebral prion infection of two independent mouse lines deficient in plasminogen resulted in either unchanged or accelerated disease progression when compared to the appropriate wild-type controls.36,37 These results do not support each other and there is no good interpretation to reconcile their incongruence. However, the discrepancy between the expectation and the actual outcomes of the previous studies can be caused by the intrinsic health problems associated with the mouse lines used for the studies,3840 which raises a question for the credibility of the model system. Some clinical phenotypes of confounding health problems in plasminogen-deficient mice overlap with the typical clinical signs of prion disease in mice. Furthermore, the drastically shortened life expectancy of plasminogen-deficient mice coincides with the incubation periods of mice inoculated with prion strains used in the previous studies. Lastly, the health problems of plasminogen-deficient mice could exacerbate the progression of the course of disease in mice inoculated with prions. Thus, either unchanged or shortened incubation periods of prion-inoculated plasminogen-deficient mice may not solely reflect the effect of plasminogen deficiency related to prion replication. Alternatively, the reason for observed outcomes from the in vivo models may be associated with the presence of functionally redundant proteins for plasminogen even under its absence. Because kringle domains of plasminogen share a well-conserved structure with those of other proteins such as tissue-type plasminogen activator, hepatocyte growth factor and apolipoprotein (a),41 it is possible to postulate that other proteins that contain kringle domains can functionally replace plasminogen. Therefore, an animal model that circumvents the drawbacks associated with the current models is needed to address the relevance of plasminogen in prion replication in vivo.Based on our study that identified plasminogen as the first cellular protein cofactor for PrPSc propagation, we anticipate an intriguing opportunity to develop future diagnosis and therapeutic intervention for prion disease. Previously, the identity of a proposed auxiliary factor was obscure so most approaches either targeted PrP isoforms or were empirical.42 Our study suggests that plasminogen is a novel target to interfere with PrPSc replication. Therefore, a variety of strategies that either deplete plasminogen or interfere with the formation of PrP-plasminogen complexes would work for development of a novel therapeutic intervention for prion disease. For instance, RNA interference of plasminogen expression, monoclonal anti-plasminogen antibody to block plasminogen binding to PrP, L-lysine that saturates the binding sites of plasminogen to PrP, and chemical agents that destabilize the structure of plasminogen are potential strategies for this purpose.In conclusion, plasminogen is a PrP ligand with the ability to stimulate PrPSc propagation. Our findings are indispensable in gaining a better understanding of the underlying mechanism for PrPSc propagation, while unveiling a new therapeutic target for prion disease.  相似文献   

20.
The accumulation of arginine in the cerebrospinal fluid and brains of patients suffering from acute neurodegenerative diseases like Alzheimer’s disease, point to defects in the metabolic pathways involving this amino acids. The deposits of neurofibrillary tangles and senile plaques perhaps as a consequence of fibrillogenesis of β-amyloid peptides has also been shown to be a hallmark in the aetiology of certain neurodegenerative diseases. Peptidylarginine deiminase (PAD II) is an enzyme that uses arginine as a substrate and we now show that PAD II not only binds with the peptides Aβ1-40, Aβ22-35, Aβ17-28, Aβ25-35 and Aβ32-35 but assists in the proteolytic degradation of these peptides with the concomitant formation of insoluble fibrils. PAD was purified in 12.5% yield and 137 fold with a specific activity of 59 μmol min?1?mg?1 from bovine brain by chromatography on diethylaminoethyl (DEAE)-Sephacel. Characterisation of the enzyme gave a pH and temperature optima of 7.5°C and 68°C, respectively, and the enzyme lost 50% activity within 38 min at this temperature. Michaelis-Menten kinetics established a V max and K m of 1.57 μmol min?1?ml?1 and 1.35 mM, respectively, with N-benzoyl arginine ethyl ester as substrate. Kinetic analysis was used to measure the affinity (K i) of the amyloid peptides to PAD with values between 1.4 and 4.6 μM. The formation of Aβ fibrils was rate limiting involving an initial lag time of about 24 h that was dependent on the concentration of the amyloid peptides. Turbidity measurements at 400 nm, Congo Red assay and Thioflavin-T staining fluorescence were used to establish the aggregation kinetics of PAD-induced fibril formation.  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