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 46 毫秒
1.
The apoptosis protection by plasminogen activator inhibitor-2(PAI-2) is dependent on a 33 amino acid fragment between helix C and D of PAI-2 which is probably due to the interaction of PAI-2 with unknown intracellular proteins. In this study, we used the fragment between helix C and D of PAI-2 as bait to screen a HeLa cell cDNA library constructed during apoptosis in a yeast two-hybrid system and retrieved a clone encoding 241 amino acids of proteasome (prosome, macropain) subunit, beta type l(PSMβ1) which plays important roles in NF-jjjjjjjjB activation. GST-pulldown experiments confirmed the interaction between PAI-2 and PSMβ1 in vitro. These data suggest that the antiapoptosis activity of PAI-2 is probably related to its interation with PSMβ1.  相似文献   

2.
Plasminogen activator inhibitor 1 (PAI-1), a member of the serpin superfamily of proteins, has been demonstrated previously to interact functionally with the glycosaminoglycan heparin (Ehrlich, H.J., Keijer, J., Preissner, K. T., Klein Gebbink, R., and Pannekoek, H. (1991) Biochemistry 30, 1021-1028). Heparin specifically enhances the rate of association between PAI-1 and thrombin about 2 orders of magnitude, whereas no effect is detected with other serine proteases (e.g. factor Xa). For the heparin-dependent serpins antithrombin III and heparin cofactor II, basic amino acid residues in and around the helix D subdomain were proposed to be involved in the binding of glycosaminoglycans. Here we employed site-directed mutagenesis of full-length PAI-1 cDNA to identify the amino acid residues that mediate heparin binding. To that end, 15 single-point mutants of PAI-1, each having individual arginyl, lysyl, or histidyl residues replaced by a neutral (alanyl) residue ("ala-scan"), and one double mutant were constructed, expressed in Escherichia coli, and purified to apparent homogeneity. The purified biologically active proteins were subjected to the following analyses: (i) heparin-dependent inhibition of thrombin; (ii) heparin-dependent formation of sodium dodecyl sulfate-stable complexes with thrombin; and (iii) binding to and elution from heparin-Sepharose. Based on the data presented, we propose that the amino acid residues Lys65, Lys69, Arg76, Lys80, and Lys88 constitute major determinants for heparin binding of PAI-1. These residues are located in and around the helix D domain and are conserved in the other heparin-dependent thrombin inhibitors, antithrombin III and heparin cofactor II.  相似文献   

3.
The interaction between type 1 plasminogen activator inhibitor (PAI-1) and fragments of vitronectin (Vn) was investigated. The PAI-1-binding domain was not destroyed when Vn was cleaved by treatment with either acid or CNBr. Acid-cleaved Vn was fractionated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and analyzed by PAI-1 ligand binding. The smallest fragment (Mr 40,000) that retained PAI-1 binding function was sequenced and shown to contain the NH2 terminus of the molecule. Further cleavage of this fragment by treatment with CNBr generated a Mr 35,000 fragment (Pro52-Asp239) that did not interact with PAI-1, and a Mr 6,000 NH2-terminal fragment (Asp1-Met51) that spanned the somatomedin B domain and contained the RGD (cell binding) sequence. The purified Mr 6,000 fragment competed with immobilized Vn for PAI-1 binding, and formed complexes with activated PAI-1. These complexes could be immunoprecipitated by antibodies to PAI-1. Synthetic peptides containing the RGD sequence had no effect on the binding of this fragment to PAI-1. These results suggest that the cell-binding and PAI-1 binding sequences of Vn occupy distinct regions in the NH2-terminal somatomedin B domain of the molecule.  相似文献   

4.
Upon activation the human bradykinin B(2) receptor (B(2)R) acts as guanine nucleotide exchange factor for the G proteins G(q/11) and G(i). Thereafter, it gets phosphorylated by G protein-coupled receptor kinases (GRKs) and recruits β-arrestins, which block further G protein activation and promote B(2)R internalization via clathrin-coated pits. As for most G protein-coupled receptors of family A, an intracellular helix 8 after transmembrane domain 7 is also predicted for the B(2)R. We show here that disruption of helix 8 in the B(2)R by either C-terminal truncation or just by mutation of a central amino acid (Lys-315) to a helix-breaking proline resulted in strong reduction of surface expression. Interestingly, this malfunction could be overcome by the addition of the membrane-permeable B(2)R antagonist JSM10292, suggesting that helix 8 has a general role for conformational stabilization that can be accounted for by an appropriate antagonist. Intriguingly, an intact helix 8, but not the C terminus with its phosphorylation sites, was indispensable for receptor sequestration and for interaction of the B(2)R with GRK2/3 and β-arrestin2 as shown by co-immunoprecipitation. Recruitment of β-arrestin1, however, required the presence of the C terminus. Taken together, our results demonstrate that helix 8 of the B(2)R plays a crucial role not only in efficient trafficking to the plasma membrane or the activation of G proteins but also for the interaction of the B(2)R with GRK2/3 and β-arrestins. Additional data obtained with chimera of B(2)R with other G protein-coupled receptors of family A suggest that helix 8 might have similar functions in other GPCRs as well.  相似文献   

5.
Recently, we found that alpha(1)-acid glycoprotein (AGP), one of the major acute-phase proteins, forms a function-stabilizing complex with plasminogen activator inhibitor 1 (PAI-1). In this study we describe the mechanism by which AGP, as well as its recombinant fragment AGP(118)(-)(201), interacts exclusively with the active form of PAI-1 and stabilizes its conformation. The binding domain of PAI-1 for AGP was initially mapped by antibodies reacting with the well-defined PAI-1 epitopes and then verified in binding assays utilizing a library of PAI-1 mutants. The latter consisted of PAI-1 molecules with individual, tandem, or grouped mutations in the epitope region of MA-55F4C12, MA-33B8, MA-33H1F7, MA-44E4, and MA-8H9D4. Solid-phase binding experiments showed that only MA-8H9D4 did not bind to the PAI-1/AGP complex, indicating that its epitope is hidden upon binding of PAI-1 to AGP. Consistently, only PAI-1 mutants with substitutions in the region of R300-D305, constituting the MA-8H9D4 epitope, showed a lack of binding or severe deficit in both the capacity and affinity of binding to AGP. These results support a location of the binding site close to the epitope within the segment connecting the regions hI with S5A. In conclusion, our present data suggest that AGP binding stabilizes the active conformation of PAI-1 by restricting the movement of beta-sheet A and thereby preventing insertion of the reactive center loop.  相似文献   

6.
Vitronectin and plasminogen activator inhibitor-1 (PAI-1) are proteins that interact in the circulatory system and pericellular region to regulate fibrinolysis, cell adhesion, and migration. The interactions between the two proteins have been attributed primarily to binding of the somatomedin B (SMB) domain, which comprises the N-terminal 44 residues of vitronectin, to the flexible joint region of PAI-1, including residues Arg-103, Met-112, and Gln-125 of PAI-1. A strategy for deletion mutagenesis that removes the SMB domain demonstrates that this mutant form of vitronectin retains PAI-1 binding (Schar, C. R., Blouse, G. E., Minor, K. M., and Peterson, C. B. (2008) J. Biol. Chem. 283, 10297-10309). In the current study, the complementary binding site on PAI-1 was mapped by testing for the ability of a battery of PAI-1 mutants to bind to the engineered vitronectin lacking the SMB domain. This approach identified a second, separate site for interaction between vitronectin and PAI-1. The binding of PAI-1 to this site was defined by a set of mutations in PAI-1 distinct from the mutations that disrupt binding to the SMB domain. Using the mutations in PAI-1 to map the second site suggested interactions between alpha-helices D and E in PAI-1 and a site in vitronectin outside of the SMB domain. The affinity of this second interaction exhibited a K(D) value approximately 100-fold higher than that of the PAI-1-somatomedin B interaction. In contrast to the PAI-1-somatomedin B binding, the second interaction had almost the same affinity for active and latent PAI-1. We hypothesize that, together, the two sites form an extended binding area that may promote assembly of higher order vitronectin-PAI-1 complexes.  相似文献   

7.
Bacillus thuringiensis (Bt) Cry8D insecticidal proteins are unique among Cry8 family proteins in terms of its insecticidal activity against adult Scarab beetles, such as Japanese beetle (Popillia japonica Newman). From the sequence homology with other Bt Cry proteins especially those active against beetles, such as Cry3Aa whose 3D structure is available, the structure of the Cry8D protein has been predicted to be a typical three-domain Cry protein type. In addition, the activation process of Cry8D in gut juice of susceptible insects is presumed to be similar to that of Cry3A (Yamaguchi et al., 2008). In this study, the activation process of Cry8Da in insect gut juice was closely examined. Japanese beetle gut juice proteases digested the 130 kDa Cry8Da protein to produce a 64 kDa protein. This 64 kDa protein was active against both adult and larval Japanese beetle and considered to be an activated toxin. N-terminal sequencing of this 64 kDa protein revealed that the Cry8Da leader sequence consisting of 63 amino acid residues from M1 to F63 was removed. As in the case of Cry3Aa, the proteases further digested the 64 kDa protein to two 8 kDa and 54 kDa fragments. N-terminal amino acid analysis of these smaller fragments indicated that the proteases digested the loop between Alpha Helix (Alpha for short) 3 and Alpha 4. This means that the 8 kDa fragment consists of Alpha 1-3 of Domain I and that the 54 kDa fragment contains the remaining Domain I and full Domain II and Domain III. Size exclusion chromatography and anion exchange chromatography could not separate these 64, 54 and 8 kDa proteins suggesting that the 54 kDa and 8 kDa fragments are still forming the toxin complex equivalent to the 64 kDa protein by size and ionic charge. The sequencing and chromatography results suggest that the gut juice proteases merely nicked the loop between Alpha 3 and Alpha 4. This nicking process appeared to be essential for receptor binding of the Cry8Da toxin. BBMV binding assay revealed that the Cry8Da toxin bound to BBMV preparations from both adult and larval Japanese beetle only after the loop was nicked. Only the 54 kDa fragment bound to the BBMV preparations but not the 64 kDa protein. Ligand blot showed that the protease activated Cry8Da toxin, presumably the 54 kDa fragment, bound to specific BBMV proteins, one or more of those would be receptor(s). The sizes and binding affinities of these Cry8Da-bound proteins of Japanese beetle BBMV differed between larvae and adults.  相似文献   

8.
Calmodulin binding to the Fas death domain. Regulation by Fas activation   总被引:3,自引:0,他引:3  
Fas (APO-1/CD95) is a cell surface receptor that initiates apoptotic pathways, and its cytoplasmic domain interacts with various molecules suggesting that Fas signaling is complex and regulated by multiple proteins. Calmodulin (CaM) is an intracellular Ca(2+)-binding protein, and it mediates many of the effects of Ca2+. Here, we demonstrate that CaM binds to Fas directly and identify the CaM-binding site on the cytoplasmic death domain (DD) of Fas. Fas binds to CaM-Sepharose and is co-immunoprecipitated with CaM. Other death receptors, such as tumor necrosis factor receptor, DR4, and DR5 do not bind to CaM. The interaction between Fas and CaM is Ca(2+)-dependent. Deletion mapping analysis with various GST-fused Fas cytoplasmic domain fragments revealed that the fragment containing helices 1, 2, and 3 of the Fas DD has the CaM-binding ability. Sequence analysis of this fragment predicted a potential CaM-binding site in helix 2 and connected loops. A valine 254 to asparagine mutation in this region, which is analogous to the identified mutant allele of Fas in lpr mice that have a deficiency in Fas-mediated apoptosis, showed reduced CaM binding. Computer modeling of the interaction between CaM and helix 2 of the Fas DD predicted that amino acids, which are important for Fas-CaM binding, and point mutations of these amino acids caused reduced Fas-CaM binding. The interaction between Fas and CaM is increased approximately 2-fold early upon Fas activation (at 30 min) and is decreased to approximately 50% of control at 2 h. These findings suggest a novel function of CaM in Fas-mediated apoptosis.  相似文献   

9.
Plasminogen activator inhibitor type 2 (PAI-2) is the only wild-type serpin that polymerizes spontaneously under physiological conditions. We show that PAI-2 loses its ability to polymerize following reduction of thiol groups, suggesting that an intramolecular disulfide bond is essential for the polymerization. A novel disulfide bond was identified between C79 (in the CD-loop) and C161 (at the bottom of helix F). Substitution mutants in which this disulfide bond was broken did not polymerize. Reactive center loop peptide insertion experiments and binding of bis-ANS to hydrophobic cavities indicate that the C79-C161 disulfide bond stabilizes PAI-2 in a polymerogenic conformation with an open A-beta-sheet. Elimination of this disulfide bond causes A-beta-sheet closure and abrogates the polymerization. The finding that cytosolic PAI-2 is mostly monomeric, whereas PAI-2 in the secretory pathway is prone to polymerize, suggests that the redox status of the cell could regulate PAI-2 polymerization. Taken together, our data suggest that the CD-loop functions as a redox-sensitive switch that converts PAI-2 between an active stable monomeric and a polymerogenic conformation, which is prone to form inactive polymers.  相似文献   

10.
Enzyme-catalyzed reformation of peptide bonds in the noncovalent fragment systems of proteins has been emerging as a convenient procedure for the semisynthesis of covalent analogs of the respective proteins. Limited proteolysis of the alpha-chain of hemoglobin S with Staphylococcus aureus V8-protease converts the chain into a fragment-complementing system by hydrolyzing the peptide bond Glu(30)-Arg(31) of the chain. Therefore, it is conceivable that semisynthesis of covalent analogs of alpha-chain could be achieved if conditions for the V8-protease catalyzed formation of peptide bonds could be established. The synthetic potential of V8-protease has been now investigated by incubating V8-protease-derived fragments of alpha-chain, namely alpha 1-30 and alpha 31-47 with the enzyme at pH 6.0 in the presence of n-propanol as the organic cosolvent. RP high performance liquid chromatography analysis showed that a new chromatographically distinct component is generated on incubation, and this has been identified as alpha 1-47 by amino acid analysis, redigestion with V8-protease (in the absence of n-propanol), and tryptic peptide mapping. Optimal conditions for the synthesis of alpha 1-47 is at pH 6.0, 4 degrees C, and 24 hr of incubation with 25% n-propanol as organic cosolvent. This stereospecific condensation of the fragments proceeded to a high level of about 50% in 24 hr. Further incubation up to 72 hr did not increase the yield of alpha 1-47, suggesting that an equilibration of synthesis and hydrolysis reactions has been attained. The demonstration of the synthetic potential of V8-protease and the fact that alpha 1-30 and alpha 31-141 interact to form a native-like complex, opens up an approach for the semisynthesis of covalent analogs of alpha-chain of hemoglobin S.  相似文献   

11.
Plasminogen activator inhibitor 2 (PAI-2) has been shown to inhibit apoptosis in transfected cells. We have investigated this phenomenon in activated human monocytes, which are a physiological source of intracellular PAI-2. Apoptosis of monocytes was rapidly induced by removal of serum, addition of hydrogen peroxide, or binding of a monoclonal antibody to Fas. Treatment of monocytes with thrombin or lipopolysaccharide (LPS) inhibited apoptosis of monocytes and also up-regulated intracellular PAI-2. Increased apoptosis was accompanied with increased activity of caspases 3 and 8. Thrombin or LPS treatment of monocytes decreased the activity of both caspases, which correlated with protection from apoptosis. The role for PAI-2 in protection of monocytes from apoptosis was studied. Monocytes were transfected with antisense oligonucleotides that blocked PAI-2 antigen, and antisense for PAI-2 had no effect on apoptosis of monocytes. No interaction was evident between PAI-2 and recombinant caspases 3 and 8 in vitro. PAI-2 was not a substrate for caspases during apoptosis of monocytes, although some cleavage of recombinant PAI-2 by caspase 3 was evident in vitro. This study shows that thrombin or LPS protected monocytes from apoptosis and that PAI-2 did not mediate this inhibitory effect.  相似文献   

12.
The Bcl-2 family of proteins plays a central role in the regulation of mitochondrial outer-membrane permeabilization, a critical step in apoptosis. Heterodimerization between the pro- and anti-apoptotic members of Bcl-2 family is a key event in this process. Anti-apoptotic proteins have high levels of expression in many cancers and they have different affinities for different pro-apoptotic proteins. Experimentally determined structures of all members of Bcl-2 proteins have remarkably similar helical fold despite poor amino acid sequence identity. Peptides representing BH3 region of pro-apoptotic proteins have been shown to bind the hydrophobic cleft of anti-apoptotic proteins and this segment is responsible in modulating the apoptotic pathways in living cells. Understanding the molecular basis of protein-protein recognition is required to develop inhibitors specific to a particular anti-apoptotic protein. We have carried out molecular dynamics simulations on the anti-apoptotic Bcl-X(L) protein in complex with three different BH3 peptides derived from pro-apoptotic Bak, Bad and Bim proteins. Each complex structure was simulated for a period of 50 ns after 2.5 ns equilibration. Analysis of the simulation results showed that in the Bcl-X(L) protein, the helix containing the BH3 region is more flexible than other helices in all three simulations. A network of strong hydrophobic interactions exists between four of the six helices and they contribute significantly to the stability of this helix bundle protein. Analysis of Bcl-X(L)-BH3 peptide interactions reveals the role of loop residues in the protein-peptide interactions in all three simulations. Bad and Bim peptides maintain strong hydrophobic and hydrophilic interactions with the helix preceding the central hydrophobic helix. Residues from this helix interact with an Arg residue in Bad and Bim peptides. This Arg residue is next to the conserved Leu residue and is replaced by Ala in Bak. Absence of these interactions and the helix propensity are likely to be the cause for Bak peptide's weaker binding affinity with the Bcl-X(L) protein. The results of this study have implications in the design of Bcl-X(L)-specific inhibitors.  相似文献   

13.
Sos and Ras-GRF are two families of guanine nucleotide exchange factors that activate Ras proteins in cells. Sos proteins are ubiquitously expressed and are activated in response to cell-surface tyrosine kinase stimulation. In contrast, Ras-GRF proteins are expressed primarily in central nervous system neurons and are activated by calcium/calmodulin binding and by phosphorylation. Although both Sos1 and Ras-GRF1 activate the Ras proteins Ha-Ras, N-Ras, and Ki-Ras, only Ras-GRF1 also activates the functionally distinct R-Ras GTPase. In this study, we determined which amino acid sequences in these exchange factors and their target GTPases are responsible for this signaling specificity difference. Analysis of chimeras and individual amino acid exchanges between Sos1 and Ras-GRF1 revealed that the critical amino acids reside within an 11-amino acid segment of their catalytic domains between the second and third structurally conserved regions (amino acids (aa) 828-838 in Sos1 and 1057-1067 in Ras-GRF1) of Ras guanine nucleotide exchange factors. In Sos1, this segment is in helix B, which is known to interact with the switch 2 region of Ha-Ras. Interestingly, a similar analysis of Ha-Ras and R-Ras chimeras did not identify the switch 2 region of Ha-Ras as encoding specificity. Instead, we found a more distal protein segment, helix 3 (aa 91-103 in Ha-Ras and 117-129 in R-Ras), which interacts instead primarily with helix K (aa 1002-1016) of Sos1. These findings suggest that specificity derives from the fact that R-Ras-specific amino acids in the region analogous to Ha-Ras helix 3 prevent a functional interaction with Sos1 indirectly, possibly by preventing an appropriate association of its switch 2 region with helix B of Sos1. Although previous studies have shown that helix B of Sos1 and helix 3 of Ha-Ras are involved in promoting nucleotide exchange on Ras proteins, this study highlights the importance of these regions in establishing signaling specificity.  相似文献   

14.
Rybp (DEDAF) has been shown to interact with DED-containing proteins and to encode pro-apoptotic functions. Here we characterize a novel interaction between Rybp and Hippi, a protein implicated in neuronal apoptosis as well as in the pathogenesis of Huntington’s disease. Rybp can synergize with Hippi to enhance Caspase 8-mediated apoptosis and also appears to be essential for Hippi-mediated apoptosis. Moreover, Rybp may mediate or regulate the interaction between Hippi and Caspase 8. Finally, Rybp and Hippi co-localize in a subset of neurons in the developing mouse brain. Together, these findings suggest that Rybp and Hippi may functionally interact in the apoptotic processes that accompany normal murine neural development.  相似文献   

15.
The PAQosome is a large complex composed of the HSP90/R2TP chaperone and a prefoldin-like module. It promotes the biogenesis of cellular machineries but it is unclear how it discriminates closely related client proteins. Among the main PAQosome clients are C/D snoRNPs and in particular their core protein NOP58. Using NOP58 mutants and proteomic experiments, we identify different assembly intermediates and show that C12ORF45, which we rename NOPCHAP1, acts as a bridge between NOP58 and PAQosome. NOPCHAP1 makes direct physical interactions with the CC-NOP domain of NOP58 and domain II of RUVBL1/2 AAA+ ATPases. Interestingly, NOPCHAP1 interaction with RUVBL1/2 is disrupted upon ATP binding. Moreover, while it robustly binds both yeast and human NOP58, it makes little interactions with NOP56 and PRPF31, two other closely related CC-NOP proteins. Expression of NOP58, but not NOP56 or PRPF31, is decreased in NOPCHAP1 KO cells. We propose that NOPCHAP1 is a client-loading PAQosome cofactor that selects NOP58 to promote box C/D snoRNP assembly.  相似文献   

16.
The signal recognition particle (SRP) is a ribonucleoprotein complex that plays a crucial role during the delivery of secretory proteins from the ribosome to the cell membrane. Among the six proteins of the eukaryotic SRP, the 72 kDa protein (SRP72) is the largest and least characterized. Polypeptides corresponding to various regions of the entire human SRP72 sequence were expressed in Escherichia coli, purified, and partially proteolyzed. Human SRP RNA bound with high affinity to a 63 amino acid residue region near the C terminus of SRP72. Mild treatment of the fragment with chymotrypsin abolished its RNA-binding activity. A conserved sequence with the consensus PDPXRWLPXXER was identified within a 56 amino acid residue RNA-binding domain. Sucrose gradient centrifugation and filter-binding analysis using mutant SRP RNAs showed that SRP72 bound to the moderately conserved portion of SRP RNA helix 5. Nine tetratricopeptide-like repeats (TPRs) poised to interact with other SRP or ribosomal proteins were predicted in the NH2-terminal region. These identifications assign two important functions to a large portion of SRP72 and demonstrate the RNA-binding capacity of the protein.  相似文献   

17.
Previously, we have demonstrated that purified Hoxc8 homeoprotein has the ability to penetrate the cellular membrane and can be transduced efficiently into COS-7 cells. Moreover, the Hoxc8 protein is able to form a complex with DNA molecules in vitro and helps the DNA be delivered intracellularly, serving as a gene delivery vehicle. Here, we further analyzed the membrane transduction activity of Hoxc8 protein and provide the evidence that the 16 amino acid (a.a.191-206, 2.23 kDa) third helix of murine Hoxc8 protein is an efficient protein transduction domain (PTD). When the 16 amino acid peptide was fused at the carboxyl terminal of enhanced green fluorescence protein (EGFP), the fusion proteins were transduced efficiently into the primary pig fetal fibroblast cells. The transduction efficiency increased in a concentration-dependent manner up to 1 μM, and appeared to plateau above a concentration of 1 μM. When tandem multimers of PTD, EGFP-PTD(2), EGFP-PTD(3), EGFP-PTD(4), and EGFP-PTD(5), were analyzed at 500 nM of concentration, the penetrating efficiency increased in a dose-dependent manner. As the number of PTDs increased, the EGFP signal also increased, although the signal maintained plateau after EGFP-PTD(3). These results indicate that the 16 amino acid third helix is the key element responsible for the membrane transduction activity of Hoxc8 proteins, and further suggest that the small peptide could serve as a therapeutic delivery vehicle for large cargo proteins.  相似文献   

18.
The carboxy-terminal alpha-helix of a nuclear receptor ligand-binding domain (LBD), helix 12, contains a critical, ligand-modulated interface for the interaction with coactivator proteins. In this study, using the example of the vitamin D receptor (VDR) and the partial antagonist ZK159222, the role of helix 12 (residues 417-427) for both antagonistic and agonistic receptor actions was investigated. Amino acid residue G423 was demonstrated to be critical for partial agonism of ZK159222, but not for the activity of the natural VDR agonist, 1alpha,25-dihydroxyvitamin D(3) (1alpha,25(OH)(2)D(3)). The amount of partial agonism of ZK159222 increased when helix 12 was truncated by the last four amino acid residues (Delta424-27) and augmented even more, when in addition helix 12 of VDR's dimerization partner, retinoid X receptor (RXR), was truncated. In contrast, the low agonism of a structural derivative of ZK159222, ZK168281, was not affected comparably, whereas other close structural relatives of ZK159222 even demonstrated the same agonistic activity as that of 1alpha,25(OH)(2)D(3). The amount of agonism of ZK159222 and ZK168281 at different variations of helix 12 correlated well with VDR's ability to complex with coactivator proteins and inversely correlated with the strength of the compound's antagonistic action on 1alpha,25(OH)(2)D(3) signalling. Molecular dynamics simulations of the LBD complexed with the two antagonists could explain their different action by demonstrating a more drastic displacement of helix 12 through ZK168281 than through ZK159222. Moreover, the modelling could indicate a kink of helix 12 at amino acid residue G423, which provides the last four amino acid residues of helix 12 with a modulatory role for the partial agonism of some VDR antagonists, such as ZK159222. In conclusion, partial agonism of a VDR antagonist is lower the more it disturbs helix 12 in taking the optimal position for coactivator interaction.  相似文献   

19.
We have delineated two different reaction mechanisms of monoclonal antibodies (mAbs), MA-8H9D4 and either MA-55F4C12 or MA-33H1F7, that convert plasminogen activator inhibitor 1 (PAI-1) to a substrate for tissue (tPA)- and urokinase plasminogen activators. MA-8H9D4 almost completely (98-99%) shifts the reaction to the substrate pathway by preventing disordering of the proteinase active site. MA-8H9D4 does not affect the rate-limiting constants (k(lim)) for the insertion of the reactive center loop cleaved by tPA (3.5 s(-1)) but decreases k(lim) for urokinase plasminogen activator from 25 to 4.0 s(-1). MA-8H9D4 does not cause deacylation of preformed PAI-1/proteinase complexes and probably acts prior to the formation of the final inhibitory complex, interfering with displacement of the acylated serine from the proteinase active site. MA-55F4C12 and MA-33H1F7 (50-80% substrate reaction) do not interfere with initial PAI-1/proteinase complex formation but retard the inhibitory pathway by decreasing k(lim) (>10-fold for tPA). Interaction of two mAbs with the same molecule of PAI-1 has been directly demonstrated for pairs MA-8H9D4/MA-55F4C12 and MA-8H9D4/MA-33H1F7 but not for MA-55F4C12/MA-33H1F7. The strong functional additivity observed for MA-8H9D4 and MA-55F4C12 demonstrates that these mAbs interact independently and affect different steps of the PAI-1 reaction mechanism.  相似文献   

20.
alpha(1)-Acid glycoprotein, one of the major acute phase proteins, was found to interact with plasminogen activator inhibitor type 1 (PAI-1) and to stabilize its inhibitory activity toward plasminogen activators. This conclusion is based on the following observations: (a) alpha(1)-acid glycoprotein was identified to bind PAI-1 by a yeast two-hybrid system. Three of 10 positive clones identified by this method to interact with PAI-1 contained almost the entire sequence of alpha(1)-acid glycoprotein; (b) this protein formed complexes with PAI-1 that could be immunoprecipitated from both the incubation mixtures and blood plasma by specific antibodies to either PAI-1 or alpha(1)-acid glycoprotein. Such complexes could be also detected by a solid phase binding assay; and (c) the real-time bimolecular interactions monitored by surface plasmon resonance indicated that the complex of alpha(1)-acid glycoprotein with PAI-1 is less stable than that formed by vitronectin with PAI-1, but in both cases, the apparent K(D) values were in the range of strong interactions (4.51 + 1.33 and 0.58 + 0.07 nm, respectively). The on rate for binding of PAI-1 to alpha(1)-glycoprotein or vitronectin differed by 2-fold, indicating much faster complex formation by vitronectin than by alpha(1)-acid glycoprotein. On the other hand, dissociation of PAI-1 bound to vitronectin was much slower than that from the alpha(1)-acid glycoprotein, as indicated by 4-fold lower k(off) values. Furthermore, the PAI-1 activity toward urokinase-type plasminogen activator and tissue-type plasminogen activator was significantly prolonged in the presence of alpha(1)-acid glycoprotein. These observations suggest that the complex of PAI-1 with alpha(1)-acid glycoprotein can play a role as an alternative reservoir of the physiologically active form of the inhibitor, particularly during inflammation or other acute phase reactions.  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