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 78 毫秒
1.
Assimilatory nitrite reductase (aNiR) reduces nitrite ions (NO$\rm{{_{2}^{-}}}$) to ammonium ions (NH$\rm{{_{4}^{+}}}$), whereas assimilatory sulfite reductase reduces sulfite (SO$\rm{{_{3}^{2-}}}$) to hydrogen sulfide (HS(-) ). Although aNiR can also reduce SO$\rm{{_{3}^{2-}}}$, its activity is much lower than when NO$\rm{{_{2}^{-}}}$ is reduced as the substrate. To increase the SO$\rm{{_{3}^{2-}}}$-reduction activity of aNiR, we performed a N226K mutation of Nii3, a representative aNiR. The resulting Nii3-N226K variant could bind non-native targets, SO$\rm{{_{3}^{2-}}}$, and HCO$\rm{{_{3}^{-}}}$, in addition to its native target, i.e., NO$\rm{{_{2}^{-}}}$. We have determined the high-resolution structure of Nii3-N226K in its apo-state and in complex with SO$\rm{{_{3}^{2-}}}$, NO$\rm{{_{2}^{-}}}$, and HCO$\rm{{_{3}^{-}}}$. This analysis revealed conformational changes of Lys226 and the adjacent Lys224 upon binding of SO$\rm{{_{3}^{2-}}}$, but not NO$\rm{{_{2}^{-}}}$. In contrast, HCO$\rm{{_{3}^{-}}}$ binding induced a conformational change at Arg179. After replacing Asn226 with a positively charged Lys, aNiR showed affinity for several anions. A comparison of all ligand-bound structures for Nii3-N226K revealed that structural changes in the active site depend on the size of the substrate.  相似文献   

2.
Tobacco expresses four isomers of assimilatory nitrite reductase (aNiR), leaf‐type (Nii1 and Nii3), and root‐type (Nii2 and Nii4). The high‐resolution crystal structures of Nii3 and Nii4, determined at 1.25 and 2.3 Å resolutions, respectively, revealed that both proteins had very similar structures. The Nii3 structure provided detailed geometries for the [4Fe–4S] cluster and the siroheme prosthetic groups. We have generated two types of Nii3 variants: one set focuses on residue Met175 (Nii3‐M175G, Nii3‐M175E, and Nii3‐M175K), a residue that is located on the substrate entrance pathway; the second set targets residue Gln448 (Nii3‐Q448K), a residue near the prosthetic groups. Comparison of the structures and kinetics of the Nii3 wild‐type (Nii3‐WT) and the Met175 variants showed that the hydrophobic side‐chain of Met175 facilitated enzyme efficiency (kcat/Km). The Nii4‐WT has Lys449 at the equivalent position of Gln448 in Nii3‐WT. The enzyme activity assay revealed that the turnover number (kcat) and Michaelis constant (Km) of Nii4‐WT were lower than those of Nii3‐WT. However, the kcat/Km of Nii4‐WT was about 1.4 times higher than that of Nii3‐WT. A comparison of the kinetics of the Nii3‐Q448K and Nii4‐K449Q variants revealed that the change in kcat/Km was brought about by the difference in Residue 448 (defined as Gln448 in Nii3 and Lys449 in Nii4). By combining detailed crystal structures with enzyme kinetics, we have proposed that Nii3 is the low‐affinity and Nii4 is the high‐affinity aNiR.  相似文献   

3.
Copper-containing nitrite reductases (Cu-NIRs) reduce nitrite to NO. Reported here are DFT (density functional theory) results on models of the Cu-NIR active site bound to nitrite and nitric oxide. The Cu-NIR active site appears to have been designed to exclude N-nitrite binding even though N-O bond cleavage would be equally facile in the N- and O-isomers. The active site also appears to force a side-on coordination of the end-product, nitric oxide. The latter feature has to rely on the sterics of the active site to destabilize, thermodynamically speaking, the Cu-NO adduct; under these conditions, the absence of N-nitrite coordination is proposed to be merely a side-effect. For the Cu(II)-NO adduct, sterical crowding appears to also favour the Cu-NO electromer over Cu(I)-NO+, helping to avoid the potentially damaging chemistry associated with an NO+ moiety. These conclusions are in reasonable agreement with previous conclusions drawn from experiment [Science 304 (2004) 867].  相似文献   

4.
PDC-109, the major heparin-binding protein of bull seminal plasma, binds to sperm choline lipids at ejaculation and modulates capacitation mediated by heparin. Affinity chromatography on heparin-Sepharose showed that polydisperse, but not monomeric, PDC-109 displayed heparin-binding capability. We sought to characterise the surface topology of the quaternary structure-dependent heparin-binding region of PDC-109 by comparing the arginine- and lysine-selective chemical modification patterns of the free and the heparin-bound protein. A combination of reversed-phase peptide mapping of endoproteinase Lys-C-digested PDC-109 derivatives and mass spectrometry was employed to identify modified and heparin-protected residues. PDC-109 contains two tandemly arranged fibronectin type II domains (a, Cys24-Cys61; b, Cys69-Cys109). The results show that six basic residues (Lys34, Arg57, Lys59, Arg64, Lys68, and Arg104) were shielded from reaction with acetic anhydride and 1,2-cyclohexanedione in heparin-bound PDC-109 oligomers. In the 1H-NMR solution structures of single fibronectin type II domains, residues topologically equivalent to PDC-109 Arg57 (Arg104) and Lys59 lay around beta-strand D on the same face of the domain. In full-length PDC-109, Arg64 and Lys68 are both located in the intervening polypeptide between domains a and b. Our data suggest possible quaternary structure arrangements of PDC-109 molecules to form a heparin-binding oligomer.  相似文献   

5.
The role of thrombin anion-binding exosite-I in the recognition and cleavage of the extracellular domain of the seven transmembrane domain thrombin receptor (PAR1) was determined using site-directed mutagenesis. Basic residues in anion-binding exosite-I (Arg35, Arg36, Arg67, Arg73, Arg75, Arg77A, Lys81, Lys109, Lys110 and Lys149E) were substituted with glutamines and the resultant recombinant mutant thrombins were used to determine kinetic parameters for the cleavage of a peptide (PAR38-60) based on the PAR1 extracellular domain. Compared with wild-type thrombin, replacement of Arg67 and Arg73 had a dramatic effect on the cleavage of PAR38-60 (k(cat)/K(m) = 1.8 x 10(6) and 4.6 x 10(6) vs 9.2 x 10(7) M(-1).s(-1)), whereas the remaining mutations of the anion-binding exosite-I of thrombin had a less pronounced effect, with k(cat)/K(m) values ranging from 3.3 x 10(7) M(-1). s(-1) (R77(a)Q) to 5.8 x 10(7) M(-1).s(-1) (K109Q). The ability of thrombin mutants to activate platelets paralleled that of PAR38-60 cleavage, whereas their ability to clot fibrinogen differed profoundly, as did their susceptibility to hirudin inhibition. Results are interpreted with respect to known interactions of thrombin with thrombomodulin, hirudin, rhodniin and heparin cofactor II. We conclude that the basic residues of anion-binding exosite-I contribute significantly to enhancing the rate of complex formation in two ways; the first (general) ensures electrostatic steering of ligands with complementary electrostatic fields, the second (specific) involves a combination of molecular contacts within the complex that is unique for each ligand.  相似文献   

6.
The mechanism of nitric oxide reduction in a ba(3)-type heme-copper oxidase has been investigated using density functional theory (B3LYP). Four possible mechanisms have been studied and free energy surfaces for the whole catalytic cycle including proton and electron transfers have been constructed by comparison to experimental data. The first nitric oxide coordinates to heme a(3) and is partly reduced having some nitroxyl anion character ((3)NO(-)), and it is thus activated toward the attack by the second N-O. In this reaction step a cyclic hyponitrous acid anhydride intermediate with the two oxygens coordinating to Cu(B) is formed. The cyclic hyponitrous acid anhydride is quite stable in a local minimum with high barriers for both the backward and forward reactions and should thus be observable experimentally. To break the N-O bond and form nitrous oxide, the hyponitrous acid anhydride must be protonated, the latter appearing to be an endergonic process. The endergonicity of the proton transfer makes the barrier of breaking the N-O bond directly after the protonation too high. It is suggested that an electron should enter the catalytic cycle at this stage in order to break the N-O bond and form N(2)O at a feasible rate. The cleavage of the N-O bond is the rate limiting step in the reaction mechanism and it has a barrier of 17.3 kcal/mol, close to the experimental value of 19.5 kcal/mol. The overall exergonicity is fitted to experimental data and is 45.6 kcal/mol.  相似文献   

7.
S H Lin  N T Yu  K Gersonde 《FEBS letters》1988,229(2):367-371
Resonance Raman spectroscopy has been employed to determine the vibrational modes of monomeric nitrosyl manganese Chironomus thummi thummi hemoglobin (CTT IV). This insect hemoglobin has no distal histidine. By applying various isotope-labeled nitric oxides (14N16O, 15N16O, 14N18O), we have identified the Mn11-NO stretching model at 628 cm-1, the Mn11-N-O bending mode at 574 cm-1 and the N-O stretching mode at 1735 cm-1. The results suggest a strong Mn11-NO bond and a weak N-O bond. The vinyl group substitution does not influence the nu (Mn11-NO), delta (Mn11-N-O) and nu (N-O) vibrations. The Mn11-NO stretching frequency is insensitive to distal histidine interactions with NO, whereas the N-O stretching frequency is sensitive. Nitric oxide also binds to Met manganese CTT IV to form an Mn111. NO complex which undergoes a slow but complete autoreduction resulting in the Mn11.NO species. In manganese meso-IX CTT IV, the Mn111. NO Mn11. NO conversion alters the intensities of the porphyrin ring modes at 342, 360, 1587 and 1598 cm-1, but shifts the frequencies at 1504 and 1633 cm-1 (in Mn111.NO) to 1497 and 1630 cm-1 (in Mn11. NO), respectively. The unshifted marker line at 1378 cm-1 reflects the fact that the pi electron densities of the porphyrin ring are the same in the two complexes.  相似文献   

8.
Rabbit myelin basic protein (BP) was subjected to partial cleavage with plasmin, and 15 cleavage products were isolated by a combination of gel filtration and ion-exchange chromatography. Their identification was achieved by amino acid analysis and tryptic peptide mapping, supplemented in some instances by carboxy-terminal analyses with carboxypeptidases A, B, and Y and amino-terminal analyses with dipeptidyl aminopeptidase I. The results showed that major plasmic cleavage sites included the Lys89-Asn90, Lys133-Ser134, and Lys153-Leu154 bonds. Cleavages also occurred at the Arg31-His32, Lys53-Arg54, and Arg25-His26 bonds, but these appeared to be less extensive. A large number of additional peptides were produced in relatively low yield. The smaller of these were isolated from heterogeneous fractions by high-voltage electrophoresis-TLC. Amino acid analysis of these peptides showed that minor cleavage sites included the Arg9-His10, Lys13-Tyr14, Lys103-Gly104, Lys137-Gly138, Lys140-Gly141, and Arg160-Ser161 bonds. In spite of a lower selectivity toward peptide bonds in BP as compared with pepsin, cathepsin D, and thrombin, plasmin has the advantage over the former proteinases in that it does not cleave at or near the Phe44-Phe45 bond. Instead it cleaves at the Arg31-His32 and Lys53-Arg54 bonds, thus preserving the entire hydrophobic sequence Ile-Leu-Asp-Ser-Ile-Gly-Arg-Phe-Phe as well as short sequences to either side.  相似文献   

9.
Apparent rate constants of tryptic hydrolysis of amide bonds containing Arg and Lys residues in beta-casein were determined by the analysis of kinetics of accumulation of 17 major peptide components revealed by high performance liquid chromatography. When studying pH influence on Arg/Lys bond cleavage preference, averaged rate constants over several Arg&bond;X and Lys&bond;X bonds were used for analysis of kinetics of wild-type trypsin, K188H, K188F, K188Y, K188W, and of K188D/D189K mutants. The pK(a1) value of 6.5 was found for all studied trypsins. For wild-type trypsin and its K188D/D189K mutant, pK(a2) was found to be 10. The lowest among studied engineered trypsins pK(a2) = 9.3 was determined for K188Y mutant. Considerable preference for the cleavage of Arg over Lys containing peptide bonds was demonstrated for all trypsins with engineered S2 site except for K188H and K188F. The comparison of individual rate constants for various bonds showed that during the hydrolysis by wild-type trypsin, the probabilities of splitting depend on secondary specificity and local hydrophobicity of amino acid residues, which are nearest to the hydrolyzed peptide bond (P2 site). The improvement of prediction of hydrolysis rates performed by the used program was achieved after considering the presence of hydrophobic neighborhood of Lys48--Ile49 and Arg202--Gly203 bonds.  相似文献   

10.
11.
Affinity labeling in situ of the Thermus thermophilus elongation factor Tu (EF-Tu) nucleotide binding site was achieved with periodate-oxidized GDP (GDPoxi) or GTP (GTPoxi) in the absence and presence of elongation factor Ts (EF-Ts). Lys52 and Lys137, both reacting with GDPoxi and GTPoxi, are located in the nucleotide binding region. In the absence of EF-Ts Lys137 and to a lesser extent Lys52 were accessible to the reaction with GTPoxi. GDPoxi reacted much more efficiently with Lys52 than with Lys137 under these conditions [Peter, M. E., Wittman-Liebold, B. & Sprinzl, M. (1988) Biochemistry 27, 9132-9138]. In the presence of EF-Ts, GDPoxi reacted more efficiently with Lys137 than with Lys52, indicating that the interaction of EF-Ts with EF-Tu.GDPoxi induces a conformation resembling that of the EF-Tu.GDPoxi complex in the absence of EF-Ts. Binding of EF-Ts to EF-Tu.GDP enhances the accessibility of the Arg59-Gly60 peptide bond of EF-Tu to trypsin cleavage. Hydrolysis of this peptide bond does not interfere with the ability of EF-Ts to bind to EF-Tu. EF-Ts is protected against trypsin cleavage by interaction with EF-Tu.GDP. High concentrations of EF-Ts did not interfere significantly with aminoacyl-tRNA.EF-Tu.GTP complex formation.  相似文献   

12.
Single-chain urokinase-type plasminogen activator (scu-PA) is converted to urokinase by hydrolysis of the Lys158-Ile159 peptide bond. Site-directed mutagenesis of Lys158 to Gly or Glu yields plasmin-resistant mutants with a 10-20-fold reduced catalytic efficiency for the activation of plasminogen [Nelles et al. (1987) J. Biol. Chem. 262, 5682-5689]. In the present study, we have further evaluated the enzymatic properties of derivatives of recombinant scu-PA (rscu-PA), produced by site-directed mutagenesis of Lys158, Ile159 or Ile160, in order to obtain additional information on the structure/function relations underlying the enzymatic properties of the single- and two-chain u-PA moieties. [Arg158]rscu-PA (rscu-PA with Lys158 substituted with Arg) appeared to be indistinguishable from wild-type rscu-PA with respect to plasminogen-activating potential (catalytic efficiency k2/Km = 0.21 mM-1 s-1 versus 0.64 mM-1 s-1), conversion to active two-chain urokinase by plasmin (k2/Km = 0.13 microM-1 s-1 versus 0.28 microM-1 s-1), as well as its specific activity (48,000 IU/mg as compared to 60,000 IU/mg) and its fibrinolytic potential in a plasma medium (50% lysis in 2 h with 2.8 micrograms/ml versus 2.1 micrograms/ml). [Pro159]rscu-PA (Ile159 substituted with Pro) and [Gly159]rscu-PA (Ile159 converted to Gly) are virtually inactive towards plasminogen (k2/Km less than 0.004 mM-1 s-1). They are however converted to inactive two-chain derivatives by plasmin following cleavage of the Arg156-Phe157 peptide bond in [Pro159]rscu-PA and of the Lys158-Gly159 peptide bond in [Gly159]rscu-PA. [Gly158,Lys160]rscu-PA (with Lys158 converted to Gly and Ile160 to Lys) has a low catalytic efficiency towards plasminogen both as a single-chain form (k2/Km = 0.012 mM-1 s-1) and as the two-chain derivative (k2/Km = 0.13 mM-1 s-1) generated by cleavage of both the Arg156-Phe157 and/or the Lys160-Gly161 peptide bonds by plasmin. These findings suggest that the enzymatic properties of rscu-PA are critically dependent on the amino acids in position 158 (requirement for Arg or Lys) and position 159 (requirement for Ile). Conversion of the basic amino acid in position 158 results in a 10-20-fold reduction of the catalytic efficiency of the single-chain molecule but yields a fully active two-chain derivative. The presence of Ile in position 159 is not only a primary determinant for the activity of the two-chain derivative, but also of the single-chain precursor. Cleavage of the Arg156-Phe157 or the Lys160-Gly161 peptide bonds by plasmin yields inactive two-chain derivatives.  相似文献   

13.
Factor B is a zymogen that carries the catalytic site of the complement alternative pathway C3 convertase. During convertase assembly, factor B associates with C3b and Mg(2+) forming a pro-convertase C3bB(Mg(2+)) that is cleaved at a single factor B site by factor D. In free factor B, a pair of salt bridges binds the Arg(234) side chain to Glu(446) and to Glu(207), forming a double latch structure that sequesters the scissile bond (between Arg(234) and Lys(235)) and minimizes its unproductive cleavage. It is unknown how the double latch is released in the pro-convertase. Here, we introduce single amino acid substitutions into factor B that preclude one or both of the Arg(234) salt bridges, and we examine their impact on several different pro-convertase complexes. Our results indicate that loss of the Arg(234)-Glu(446) salt bridge partially stabilizes C3bB(Mg(2+)). Loss of the Arg(234)-Glu(207) salt bridge has lesser effects. We propose that when factor B first associates with C3b, it bears two intact Arg(234) salt bridges. The complex rapidly dissociates unless the Arg(234)-Glu(446) salt bridge is released whereupon conformational changes occur that activate the metal ion-dependent adhesion site and partially stabilize the complex. The remaining salt bridge is then released, exposing the scissile bond and permitting factor D cleavage.  相似文献   

14.
Phosphonatase functions in the 2-aminoethylphosphonate (AEP) degradation pathway of bacteria, catalyzing the hydrolysis of the C-P bond in phosphonoacetaldehyde (Pald) via formation of a bi-covalent Lys53ethylenamine/Asp12 aspartylphosphate intermediate. Because phosphonatase is a member of the haloacid dehalogenase superfamily, a family predominantly comprised of phosphatases, the question arises as to how this new catalytic activity evolved. The source of general acid-base catalysis for Schiff-base formation and aspartylphosphate hydrolysis was probed using pH-rate profile analysis of active-site mutants and X-ray crystallographic analysis of modified forms of the enzyme. The 2.9 A X-ray crystal structure of the mutant Lys53Arg complexed with Mg2+ and phosphate shows that the equilibrium between the open and the closed conformation is disrupted, favoring the open conformation. Thus, proton dissociation from the cap domain Lys53 is required for cap domain-core domain closure. The likely recipient of the Lys53 proton is a water-His56 pair that serves to relay the proton to the carbonyl oxygen of the phosphonoacetaldehyde (Pald) substrate upon addition of the Lys53. The pH-rate profile analysis of active-site mutants was carried out to test this proposal. The proximal core domain residues Cys22 and Tyr128 were ruled out, and the role of cap domain His56 was supported by the results. The X-ray crystallographic structure of wild-type phosphonatase reduced with NaBH4 in the presence of Pald was determined at 2.4A resolution to reveal N epsilon-ethyl-Lys53 juxtaposed with a sulfate ligand bound in the phosphate site. The position of the C2 of the N-ethyl group in this structure is consistent with the hypothesis that the cap domain N epsilon-ethylenamine-Lys53 functions as a general base in the hydrolysis of the aspartylphosphate bi-covalent enzyme intermediate. Because the enzyme residues proposed to play a key role in P-C bond cleavage are localized on the cap domain, this domain appears to have evolved to support the diversification of the HAD phosphatase core domain for catalysis of hydrolytic P-C bond cleavage.  相似文献   

15.
Guangcai Ma  Wenyou Zhu  Yongjun Liu 《Proteins》2016,84(11):1606-1615
Pectate lyase utilizes the anti‐β‐elimination chemistry to catalyze the cleavage of α‐1,4 glycosidic bond between D‐galacturonate regions during the degradation of plant polysaccharide pectin. We report here detailed mechanistic studies of the Bacillus subtilis pectate lyase (BsPel) using QM/MM calculations. It was found that the residue Arg279 serves as the catalytic base to abstract the α‐proton from C52 atom of substrate Ada2 subsite, forming an unstable carbanion intermediate. The glycosidic bond of this intermediate is scissile to generate the 4,5‐unsaturated digalacturonate product and a negatively charged β‐leaving group. Two active site residues (Lys247 and Arg279) and two Ca2+ ions (Ca2 and Ca3) form hydrogen‐bonding and coordination interactions with C52? COO? of Ada2, respectively, which facilitate the proton abstraction and stabilize the generated carbanion intermediates. Arg284 is not the potential proton donor to saturate the leaving group. Actually, the proton source of leaving group is the solvent water molecule rather than any active site acidic residues. In addition, the calculation results suggest that careful selections of QM‐ and Active‐regions are essential to accurately explore the enzymatic reactions. Proteins 2016; 84:1606–1615. © 2016 Wiley Periodicals, Inc.  相似文献   

16.
17.
Activation of factor VIII by factor Xa is followed by proteolytic inactivation resulting from cleavage within the A1 subunit (residues 1-372) of factor VIIIa. Factor Xa attacks two sites in A1, Arg(336), which precedes the highly acidic C-terminal region, and a recently identified site at Lys(36). By using isolated A1 subunit as substrate for proteolysis, production of the terminal fragment, A1(37-336), was shown to proceed via two pathways identified by the intermediates A1(1-336) and A1(37-372) and generated by initial cleavage at Arg(336) and Lys(36), respectively. Appearance of the terminal product by the former pathway was 7-8-fold slower than the product obtained by the latter pathway. The isolated A1 subunit was cleaved slowly, independent of the presence of phospholipid. The A1/A3-C1-C2 dimer demonstrated an approximately 3-fold increased cleavage rate constant, and inclusion of phospholipid further enhanced this value by approximately 2-fold. Although association of A1 or A1(37-372) with A3-C1-C2 enhanced the rate of cleavage at Arg(336), inclusion of A3-C1-C2 did not affect the cleavage at Lys(36) in A1(1-336). A synthetic peptide 337-372 blocked the cleavage at Lys(36) (IC(50) = 230 microm) while showing little if any effect on cleavage at Arg(336). Proteolysis at Lys(36), and to a lesser extent Arg(336), was inhibited in a dose-dependent manner by heparin. These results suggest that inactivating cleavages catalyzed by factor Xa at Lys(36) and Arg(336) are regulated in part by the A3-C1-C2 subunit. Furthermore, cleavage at Lys(36) appears to be selectively modulated by the C-terminal acidic region of A1, a region that may interact with factor Xa via its heparin-binding exosite.  相似文献   

18.
According to the X-ray structure analysis of an EcoRI-oligodeoxynucleotide complex [McClarin et al. (1986) Science 234, 1526], sequence specificity is mediated by 12 hydrogen bonds, 6 from each of the two identical subunits of the dimeric enzyme to the recognition site -GAATTC-: Arg200 forms two hydrogen bonds with guanine, while Glu144 and Arg145 form four hydrogen bonds to adjacent adenine residues. Changing the hydrogen-bonding potential at the recognition site without perturbing the rest of the interface should lead to the recognition of degenerate sequences [Rosenberg et al. (1987) in Protein Engineering (Oxender, D. L., & Fox, C. F., Eds.) pp 237-250, Liss, New York]. We have shown previously that replacing Glu144 by Gln and Arg145 by Lys affects the activity of the enzyme, not, however, its specificity [Wolfes et al. (1986) Nucleic Acids Res. 14, 9063]. We show now that also the mutation of Arg200 to Lys, the double mutation Glu144Arg145 to GlnLys, and the triple mutation Glu144Arg145Arg200 to GlnLysLys do not lead to a detectable degeneracy of the specificity of cleavage by EcoRI but significantly impair the catalytic activity of this enzyme. A detailed analysis of the steady-state kinetics of cleavage of pUC8 DNA and a tridecadeoxynucleotide substrate demonstrates that the reduction in activity for all DNA binding site mutants investigated so far is mainly due to a decrease in kcat, with the exception of the Arg200 to Lys mutant, which is only impaired in its KM.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Plasmin mainly cleaved the Arg5-Ser6 bond of Arg-Val-Leu-Pro-Arg-interleukin-8 (AVLPR-IL-8) produced by human dermal fibroblasts, which resulted in the conversion of AVLPR-IL-8 to IL-8 and the inactive pentapeptide, though a minor cleavage of AVLPR-IL-8 by plasmin at Lys8-Glu9 bond occurred.  相似文献   

20.
Plasmin not only functions as a key enzyme in the fibrinolytic system but also directly inactivates factor VIII and other clotting factors such as factor V. However, the mechanisms of plasmin-catalyzed factor VIII inactivation are poorly understood. In this study, levels of factor VIII activity increased approximately 2-fold within 3 min in the presence of plasmin, and subsequently decreased to undetectable levels within 45 min. This time-dependent reaction was not affected by von Willebrand factor and phospholipid. The rate constant of plasmin-catalyzed factor VIIIa inactivation was approximately 12- and approximately 3.7-fold greater than those mediated by factor Xa and activated protein C, respectively. SDS-PAGE analysis showed that plasmin cleaved the heavy chain of factor VIII into two terminal products, A1(37-336) and A2 subunits, by limited proteolysis at Lys(36), Arg(336), Arg(372), and Arg(740). The 80-kDa light chain was converted into a 67-kDa subunit by cleavage at Arg(1689) and Arg(1721), identical to the pattern induced by factor Xa. Plasmin-catalyzed cleavage at Arg(336) proceeded faster than that at Arg(372), in contrast to proteolysis by factor Xa. Furthermore, breakdown was faster than that in the presence of activated protein C, consistent with rapid inactivation of factor VIII. The cleavages at Arg(336) and Lys(36) occurred rapidly in the presence of A2 and A3-C1-C2 subunits, respectively. These results strongly indicated that cleavage at Arg(336) was a central mechanism of plasmin-catalyzed factor VIII inactivation. Furthermore, the cleavages at Arg(336) and Lys(36) appeared to be selectively regulated by the A2 and A3-C1-C2 domains, respectively, interacting with plasmin.  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