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.
Plasminogen activator inhibitor-1 (PAI-1) is the primary inhibitor of the plasminogen activators (PAs), tissue-type plasminogen activator (tPA), and urokinase-type plasminogen activator (uPA). A library of PAI-1 mutants containing substitutions at the P1 and P1' positions was screened for functional activity against tPA and thrombin. Several PAI-1 variants that were inactive against uPA in a previous study (Sherman, P. M., Lawrence, D. A., Yang, A. Y., Vandenberg, E. T., Paielli, D., Olson, S. T., Shore, J. D., and Ginsburg, D. (1992) J. Biol. Chem. 267, 7588-7595) had significant inhibitory activity toward tPA. This set of tPA-specific PAI-1 mutants contained a wide range of amino acid substitutions at P1 including Asn, Gln, His, Ser, Thr, Leu, Met, and all the aromatic amino acids. This group of mutants also demonstrated a spectrum of substitutions at P1'. Kinetic analyses of selected variants identified P1Tyr and P1His as the most efficient tPA-specific inhibitors, with second-order rate constants (ki) of 4.0 x 10(5) M-1s-1 and 3.6 x 10(5) M-1s-1, respectively. Additional PA-specific PAI-1 variants containing substitutions at P3 through P1' were constructed. P3Tyr-P2Ser-P1Lys-P1'Trp and P3Tyr-P2Ser-P1Tyr-P1'Met had ki values of 1.7 x 10(6) M-1s-1 and 2.5 x 10(6) M-1s-1 against tPA, respectively, but both were inactive against uPA. In contrast, P2Arg-P1Lys-P1'Ala inhibited uPA 74-fold more rapidly than tPA. The mutant PAI-1 library was also screened for inhibitory activity toward thrombin in the presence and absence of the cofactor heparin. While wild-type PAI-1 and several P1Arg variants inhibited thrombin in the absence of heparin, a number of variants were thrombin inhibitors only in the presence of heparin. These results demonstrate the importance of the reactive center residues in determining PAI-1 target specificity and suggest that second sites of interaction between inhibitors and proteases can also contribute to target specificity. Finally, the PA-specific mutants described here should provide novel reagents for dissecting the physiological role of PAI-1 both in vitro and in vivo.  相似文献   

2.
Tyrosine 34 is a prominent and conserved residue in the active site of the manganese superoxide dismutases in organisms from bacteria to man. We have prepared the mutant containing the replacement Tyr 34 --> Phe (Y34F) in human manganese superoxide dismutase (hMnSOD) and crystallized it in two different crystal forms, orthorhombic and hexagonal. Crystal structures of hMnSOD Y34F have been solved to 1.9 A resolution in a hexagonal crystal form, denoted as Y34Fhex, and to 2.2 A resolution in an orthorhombic crystal form, denoted as Y34Fortho. Both crystal forms give structures that are closely superimposable with that of wild-type hMnSOD, with the phenyl rings of Tyr 34 in the wild type and Phe 34 in the mutant very similar in orientation. Therefore, in Y34F, a hydrogen-bonded relay that links the metal-bound hydroxyl to ordered solvent (Mn-OH to Gln 143 to Tyr 34 to H2O to His 30) is broken. Surprisingly, the loss of the Tyr 34 hydrogen bonds resulted in large increases in stability (measured by Tm), suggesting that the Tyr 34 hydroxyl does not play a role in stabilizing active-site architecture. The functional role of the side chain hydroxyl of Tyr 34 can be evaluated by comparison of the Y34F mutant with the wild-type hMnSOD. Both wild-type and Y34F had kcat/Km near 10(9) M-1 s-1, close to diffusion-controlled; however, Y34F showed kcat for maximal catalysis smaller by 10-fold than the wild type. In addition, the mutant Y34F was more susceptible to product inhibition by peroxide than the wild-type enzyme. This activity profile and the breaking of the hydrogen-bonding chain at the active site caused by the replacement Tyr 34 --> Phe suggest that Tyr 34 is a proton donor for O2* - reduction to H2O2 or is involved indirectly by orienting solvent or other residues for proton transfer. Up to 100 mM buffers in solution failed to enhance catalysis by either Y34F or the wild-type hMnSOD, suggesting that protonation from solution cannot enhance the release of the inhibiting bound peroxide ion, likely reflecting the enclosure of the active site by conserved residues as shown by the X-ray structures. The increased thermostability of the mutant Y34F and equal diffusion-controlled activity of Y34F and wild-type enzymes with normal superoxide levels suggest that evolutionary conservation of active-site residues in metalloenzymes reflects constraints from extreme rather than average cellular conditions. This new hypothesis that extreme rather than normal substrate concentrations are a powerful constraint on residue conservation may apply most strongly to enzyme defenses where the ability to meet extreme conditions directly affects cell survival.  相似文献   

3.
A recent study indicated that negatively charged substances such as heparin and dextran sulfate accelerate thrombin activation of coagulation factor XI by a template mechanism. Because the serine proteinase of the natural anticoagulant pathway, activated protein C, can bind heparin, it was reasonable to think that these compounds may also bind protein C (PC) and accelerate its activation by thrombin or other heparin binding plasma serine proteinases by a similar mechanism. To test this, PC activation by thrombin and factor Xa (fXa) was studied in the presence of these polysaccharides. With thrombin in the absence of thrombomodulin (TM), these polysaccharides markedly reduced the Km for PC and Gla-domainless PC (GDPC) activation in the presence of Ca2+. With TM containing chondroitin sulfate, heparin did not influence PC activation by thrombin, but with TM lacking chondroitin sulfate, the characteristic high-affinity PC interaction at low Ca2+ (approximately 50 to 100 micromol/L) was largely eliminated by heparin. In EDTA, heparin enhanced thrombin activation of GDPC by reducing the Km, but it inhibited PC activation by increasing the Km. PC activation in EDTA was insensitive to the presence of heparin if the exosite 2 mutant, R93,97,101A thrombin, was used for activation. These results suggest that, when the Gla-domain of PC is not fully stabilized by Ca2+, it interacts with the anion binding exosite 2 of thrombin and that heparin binding to this site prevents this interaction. Additional studies indicated that, in the presence of phospholipid vesicles, heparin and dextran sulfate dramatically accelerate PC activation by fXa by also reducing the Km. Interestingly, on phospholipids containing 40% phosphatidylethanolamine, the activation rate of near physiological PC concentrations ( approximately 80 nmol/L) by fXa in the presence of dextran sulfate was nearly comparable to that observed by the thrombin-TM complex. The biochemical and potential therapeutical ramifications of these findings are discussed.  相似文献   

4.
The enzymatic properties and the three-dimensional structure of spinach glycolate oxidase which has the active-site Tyr129 replaced by Phe (Y129F glycolate oxidase) has been studied. The structure of the mutant is unperturbed which facilitates interpretation of the biochemical data. Y129F glycolate oxidase has an absorbance spectrum with maxima at 364 and 450 nm (epsilon max = 11400 M-1 cm-1). The spectrum indicates that the flavin is in its normal protonated form, i.e. the Y129F mutant does not lower the pKa of the N(3) of oxidized flavin as does the wild-type enzyme [Macheroux, P., Massey, V., Thiele, D. J., and Volokita, M. (1991) Biochemistry 30, 4612-4619]. This was confirmed by a pH titration of Y129F glycolate oxidase which showed that the pKa is above pH 9. In contrast to wild-type glycolate oxidase, oxalate does not perturb the absorbance spectrum of Y129F glycolate oxidase. Moreover oxalate does not inhibit the enzymatic activity of the mutant enzyme. Typical features of wild-type glycolate oxidase that are related to a positively charged lysine side chain near the flavin N(1)-C(2 = O), such as stabilization of the anionic flavin semiquinone and formation of tight N(5)-sulfite adducts, are all conserved in the Y129F mutant protein. Y129F glycolate oxidase exhibited about 3.5% of the wild-type activity. The lower turnover number for the mutant of 0.74 s-1 versus 20 s-1 for the wild-type enzyme amounts to an increase of the energy of the transition state of about 7.8 kJ/mol. Steady-state analysis gave Km values of 1.5 mM and 7 microM for glycolate and oxygen, respectively. The Km for glycolate is slightly higher than that found for wild-type glycolate oxidase (1 mM) whereas the Km for oxygen is much lower. As was the case for wild-type glycolate oxidase, reduction was found to be the rate-limiting step in catalysis, with a rate of 0.63 s-1. The kinetic properties of Y129F glycolate oxidase provide evidence that the main function of the hydroxyl group of Tyr129 is the stabilization of the transition state.  相似文献   

5.
Tyr114 and Tyr197 are highly conserved residues in the active site of human glutathione reductase, Tyr114 in the glutathione disulfide (GSSG) binding site and Tyr197 in the NADPH site. Mutation of either residue has profound effects on catalysis. Y197S and Y114L have 17% and 14% the activity of the wild-type enzyme, respectively. Mutation of Tyr197, in the NADPH site, leads to a decrease in Km for GSSG, and mutation of Tyr114, in the GSSG site, leads to a decrease in Km for NADPH. This behavior is predicted for enzymes operating by a ping-pong mechanism where both half-reactions partially limit turnover. Titration of the wild-type enzyme or Y114L with NADPH proceeds in two phases, Eox to EH2 and EH2 to EH2-NADPH. In contrast, Y197S reacts monophasically, showing that excess NADPH fails to enhance the absorbance of the thiolate-FAD charge-transfer complex, the predominant EH2 form of glutathione reductase. The reductive half-reactions of the wild-type enzyme and of Y114L are similar; FAD reduction is fast (approximately 500 s-1 at 4 degreesC) and thiolate-FAD charge-transfer complex formation has a rate of 100 s-1. In Y197S, these rates are only 78 and 5 s-1, respectively. The oxidative half-reaction, the rate of reoxidation of EH2 by GSSG, of the wild-type enzyme is approximately 4-fold faster than that of Y114L. These results are consistent with Tyr197 serving as a gate in the binding of NADPH, and they indicate that Tyr114 assists the acid catalyst His467'.  相似文献   

6.
Two catalytically inactive mutant forms of 1-aminocyclopropane-1-carboxylate (ACC) synthase, Y85A and K273A, were mixed in low concentrations of guanidine hydrochloride (GdnHCl). About 15% of the wild-type activity was recovered (theoretical 25% for a binomial distribution), proving that the functional unit of the enzyme is a dimer, or theoretically, a higher order oligomer. The enzyme catalyzes the conversion of S-adenosyl-L-methionine (SAM) to ACC. The value of kcat/KM is 1.2 x 10(6) M-1 s-1 at pH 8.3. Viscosity variation experiments with glycerol and sucrose as viscosogenic reagents showed that this reaction is nearly 100% diffusion controlled. The sensitivity to viscosity for the corresponding reaction of the less reactive Y233F mutant is much reduced, thus the latter reaction serves as a control for that of the wild-type enzyme. The kcat/KM vs pH profile for wild-type enzyme exhibits pKa values of 7.5 and 8.9. The former is assigned to the pKa of the alpha-amino group of SAM, while the latter corresponds to the independently determined spectrophotometric pKa of the internal aldimine. The kcat vs pH profile exhibits similar pKas, which means that the above pKa values are not perturbed in the Michaelis complex. The phenolic hydroxyl group of Tyr233 forms a hydrogen bond to the 3'-O- of PLP. The spectral and kinetic pKa (kcat/KM) values of the Y233F mutant are not identical (spectral 10.2, kinetic 8.7). A model that accounts quantitatively for these data posits two parallel pathways to the external aldimine for this mutant, the minor one has the alpha-amino group free base form of SAM reacting with the protonated imine form of the enzyme with kcat/KM approximately 6.0 x 10(3) M-1 s-1, while the major pathway involves reaction of the aldehyde form of PLP with SAM with kcat/KM approximately 7.0 x 10(5) M-1 s-1. The spectral pKa is defined only by the less reactive species.  相似文献   

7.
Bacterial Delta5-3-ketosteroid isomerase (KSI) catalyzes a stereospecific isomerization of steroid substrates at an extremely fast rate, overcoming a large disparity of pKa values between a catalytic residue and its target. The crystal structures of KSI from Pseudomonas putida and of the enzyme in complex with equilenin, an analogue of the reaction intermediate, have been determined at 1.9 and 2.5 A resolution, respectively. The structures reveal that the side chains of Tyr14 and Asp99 (a newly identified catalytic residue) form hydrogen bonds directly with the oxyanion of the bound inhibitor in a completely apolar milieu of the active site. No water molecule is found at the active site, and the access of bulk solvent is blocked by a layer of apolar residues. Asp99 is surrounded by six apolar residues, and consequently, its pKa appears to be elevated as high as 9.5 to be consistent with early studies. No interaction was found between the bound inhibitor and the residue 101 (phenylalanine in Pseudomonas testosteroni and methionine in P. putida KSI) which was suggested to contribute significantly to the rate enhancement based on mutational analysis. This observation excludes the residue 101 as a potential catalytic residue and requires that the rate enhancement should be explained solely by Tyr14 and Asp99. Kinetic analyses of Y14F and D99L mutant enzymes demonstrate that Tyr14 contributes much more significantly to the rate enhancement than Asp99. Previous studies and the structural analysis strongly suggest that the low-barrier hydrogen bond of Tyr14 (>7.1 kcal/mol), along with a moderate strength hydrogen bond of Asp99 ( approximately 4 kcal/mol), accounts for the required energy of 11 kcal/mol for the transition-state stabilization.  相似文献   

8.
The influence of the substrate P1' position on the specificity of two zinc matrix metalloproteases, membrane type-1 matrix metalloprotease (MT1-MMP) and stromelysin-3 (ST3), was evaluated by synthesizing a series of fluorogenic substrates of general formula dansyl-Pro-Leu-Ala-Xaa-Trp-Ala-Arg-NH2, where Xaa in the P1' position represents unusual amino acids containing either long arylalkyl or alkyl side chains. Our data demonstrate that both MT1-MMP and ST3 cleave substrates containing in their P1' position unusual amino acids with extremely long side chains more efficiently than the corresponding substrates with natural phenylalanine or leucine amino acids. In this series of substrates, the replacement of leucine by S-para-methoxybenzyl cysteine increased the kcat/Km ratio by a factor of 37 for MT1-MMP and 9 for ST3. The substrate with a S-para-methoxybenzyl cysteine residue in the P1' position displayed a kcat/Km value of 1.59 10(6) M-1 s-1 and 1.67 10(4) M-1 s-1, when assayed with MT1-MMP and ST3, respectively. This substrate is thus one of the most rapidly hydrolyzed substrates so far reported for matrixins, and is the first synthetic peptide efficiently cleaved by ST3. These unexpected results for these two matrixins suggest that extracellular proteins may be cleaved by matrixins at sites containing amino acids with unusual long side chains, like those generated in vivo by some post-translational modifications.  相似文献   

9.
Kex2 protease from the yeast Saccharomyces cerevisiae is the prototype for a family of eukaryotic proprotein processing proteases. To clarify understanding of the interactions responsible for substrate recognition in this family of enzymes, we have carried out a systematic examination of Kex2 substrate specificity using internally consistent sets of substrates having substitutions at only one or two positions. We examined Kex2 sequence recognition for residues at P3, P2, and P1 using two types of fluorogenic peptide substrates, peptidyl-methylcoumarinamides and internally quenched substrates in which cleavage occurs at an actual peptide bond. Kinetic analysis of the two sets of substrates gave comparable data on specificity at these three positions. For the best substrate sequences, high catalytic constants (kCM/KM) of (2-5) x 10(7) M-1 s-1 were seen for cleavage of both peptidyl-methylcoumarinamides and peptide bonds. While no evidence for positive interactions with the P3 residue emerged, Kex2 was found to discriminate against at least one residue Asp. at this position. Specificity at P2 was shown to rely primarily on recognition of a positive charge, although steric constraints on the P2 side chain were also apparent. Kex2 was demonstrated to be exquisitely selective for Arg at P1. Substitutions with similar charge (Lys, ornithine) or similar hydrogen-bonding capability (citrulline) do not confer efficient catalysis. Comparison of otherwise identical substrates having either Arg or citrulline at P1 showed that the positive charge of the Arg guanidinium group stabilizes the transition state by approximately 6.8 kcal/mol.  相似文献   

10.
DoxA is a cytochrome P-450 monooxygenase involved in the late stages of daunorubicin and doxorubicin biosynthesis that has a broad substrate specificity for anthracycline glycone substrates. Recombinant DoxA was purified to homogeneity from Streptomyces lividans transformed with a plasmid containing the Streptomyces sp. strain C5 doxA gene under the control of the strong SnpR-activated snpA promoter. The purified enzyme was a monomeric, soluble protein with an apparent Mr of 47,000. Purified DoxA catalyzed the 13-hydroxylation of 13-deoxydaunorubicin, the 13-oxidation of 13-dihydrocarminomycin and 13-dihydrodaunorubicin, and the 14-hydroxylation of daunorubicin. The pH optimum for heme activation was pH 7.5, and the temperature optimum was 30 degreesC. The kcat/Km values for the oxidation of anthracycline substrates by purified DoxA, incubated with appropriate electron-donating components, were as follows: for 13-deoxydaunorubicin, 22,000 M-1 x s-1; for 13-dihydrodaunorubicin, 14,000 M-1 x s-1; for 13-dihydrocarminomycin, 280 M-1 x s-1; and for daunorubicin, 130 M-1 x s-1. Our results indicate that the conversion of daunorubicin to doxorubicin by this enzyme is not a favored reaction and that the main anthracycline flux through the late steps of the daunorubicin biosynthetic pathway catalyzed by DoxA is likely directed through the 4-O-methyl series of anthracyclines.  相似文献   

11.
The Tyr alpha 42-Asp beta 99 intersubunit H-bond stabilizes the T quaternary structure in hemoglobin (Hb) tetramers. We had proposed that Tyr alpha 42 acts as an acceptor in this H-bond, because the tyrosine Y8a/8b and Y7a' UVRR (ultraviolet resonance Raman) bands shift in directions opposite to those expected if tyrosine is an H-bond donor. If Asp beta 99 is the H-bond donor, then it must be protonated in the T state, and would be a previously unrecognized contributor to the Bohr effect. This implication was strengthened by the discovery that an R-minus-T difference FTIR (Fourier transform infrared) band at 1693 cm-1, which might be a signal from protonated carboxylate, is missing in Hb Kempsey, a mutant in which Asp beta 99 is replaced by Asn. However, we now find that this FTIR signal is insensitive to 13C-labeling of the aspartate residues in Hb, and cannot arise from protonated Asp beta 99. There are no other difference signals in the 1700 cm-1 region at a sensitivity of one COOH group. We conclude that Asp beta 99 is not protonated, and that the anomalous UVRR shifts must arise from compensating polarization of the Tyr alpha 42 OH. Candidates for this compensation are the H-bond donated by the Asp beta 94 backbone NH, and the nearby positive charge of Arg beta 40.  相似文献   

12.
Meizothrombin and meizothrombin(desF1) are intermediates formed during the conversion of prothrombin to thrombin by factor Xa, factor Va, phospholipids, and Ca2+ (prothrombinase). These intermediates are active toward synthetic peptide substrates but have limited ability to interact with platelets or macromolecular substrates such as fibrinogen. Meizothrombin and meizothrombin(desF1) activate protein C, however, and may exert primarily an anticoagulant effect. In this study, we investigated the inhibition of meizothrombin and meizothrombin(desF1) by two glycosaminoglycan-dependent protease inhibitors, heparin cofactor II (HCII) and antithrombin (AT). Purified recombinant meizothrombin and meizothrombin(desF1) were inhibited by HCII in the presence of dermatan sulfate with maximal second-order rate constants of 8 x 10(6) M-1.min-1 and 1.8 x 10(7) M-1.min-1, respectively, but were inhibited less than one-tenth as fast by AT in the presence of heparin. Similarly, the products of the prothrombinase reaction were inhibited in situ more effectively by HCII than by AT. When HCII and dermatan sulfate were present continuously during the prothrombinase reaction, meizothrombin was trapped as a sodium dodecyl sulfate-stable complex with HCII and no amidolytic activity could be detected with a thrombin substrate. Our findings indicate that HCII is an effective inhibitor of meizothrombin and meizothrombin(desF1) and, therefore, might regulate the anticoagulant activity of these proteases.  相似文献   

13.
The transmembrane PTPase HPTP beta differs from its related family members in having a single rather than a tandemly duplicated cytosolic catalytic domain. We have expressed the 354-amino acid, 41-kDa human PTP beta catalytic fragment in Escherichia coli, purified it, and assessed catalytic specificity with a series of pY peptides. HPTP beta shows distinctions from the related LAR PTPase and T cell CD45 PTPase domains: it recognizes phosphotyrosyl peptides of 9-11 residues from lck, src, and PLC gamma with Km values of 2, 4, and 1 microM, some 40-200-fold lower than the other two PTPases. With kcat values of 30-205 s-1, the catalytic efficiency, kcat/Km, of the HPTP beta 41-kDa catalytic domain is very high, up to 5.7 x 10(7) M-1 s-1. The peptides corresponding to PLC gamma (766-776) and EGFR (1,167-1,177) phosphorylation sites were used for structural variation to assess pY sequence context recognition by HPTP beta catalytic domain. While exchange of the alanine residue at the +2 position of the PLC gamma (Km of 1 microM) peptide to lysine or aspartic acid showed little or no effect on substrate affinity, replacement by arginine increased the Km 35-fold. Similarly, the high Km value of the EGFR pY peptide (Km of 104 microM) derives largely from the arginine residue at the +2 position of the peptide, since arginine to alanine single mutation at the -2 position of the EGFR peptide decreased the Km value 34-fold to 3 microM. Three thiophosphotyrosyl peptides have been prepared and act as substrates and competitive inhibitors of these PTPase catalytic domains.  相似文献   

14.
The hypoxanthine-guanine-xanthine phosphoribosyltransferase (HGXPRTase) from Tritrichomonas foetus has been proven to be a target for potential anti-tritrichomonial chemotherapy. Using a structure-based approach, the base-binding region of the active site of this enzyme, which confers unique purine base specificity, was characterized using site-directed mutagenesis. Determining the roles of different active-site residues in purine specificity would form the basis for designing specific inhibitors toward the parasitic enzyme. A D163N mutant converts the HGXPRTase into a HGPRTase, which no longer recognizes xanthine as a substrate, whereas specificities toward guanine and hypoxanthine are unaffected. Apparently, the side-chain carboxyl of Asp163 forms a hydrogen bond through a water molecule with the C2-carbonyl of xanthine, which constitutes the critical force enabling the enzyme to recognize xanthine as a substrate. Mutations of Arg155, which orients and stacks the neighboring Tyr156 onto the bound purine base by forming a salt bridge between itself and Glu11, result in drastic increases in the Kms for GMP and XMP (but not IMP). This change leads to increased kcats for the forward reactions with guanine and xanthine as substrates without affecting the conversion of hypoxanthine to IMP. Thus, the apparent dislocation of Tyr156, resulted from mutations of Arg155, bring little effect on the hydrophobic interactions between Tyr156 and the purine ring. But the forces involved in recognizing the exocyclic C2-substituents of the purine ring, which involve the Tyr156 hydroxyl, Ile157 backbone carbonyl, and Asp163 side-chain carboxyl, may be weakened by the shifted conformation of the peptide backbone resulted from loss of the Glu11-Arg155 salt bridge. The conserved Lys134 was proven to be the primary determinant in conferring the specificity of the enzyme toward 6-oxopurines. By substituting the lysine residue for a serine, which can potentially hydrogen bond to either an amino or an oxo-group, we have successfully augmented the purine specificity of the enzyme. The K134S mutant recognizes adenine in addition to hypoxanthine, guanine, and xanthine as its substrates. Adenine and hypoxanthine are equivalent substrates for the mutant enzyme with similar Kms of 34.6 and 38.0 microM, respectively. The catalysis of an adenine phosphoribosyltransferase reaction by this mutant enzyme was further demonstrated by the competitive inhibition of AMP with an estimated Kis of 25.4 microM against alpha-D-5-phosphoribosyl-pyrophosphate (PRPP) in converting hypoxanthine to IMP. We have thus succeeded in using site-directed mutagenesis to convert T. foetusHGXPRTase into either a HGPRTase or a genuine AHGXPRTase.  相似文献   

15.
BACKGROUND: The serine protease prostate-specific antigen (PSA) is a useful clinical marker for prostatic malignancy. PSA is a member of the kallikrein subgroup of the (chymo)trypsin serine protease family, but differs from the prototypical member of this subgroup, tissue kallikrein, in possessing a specificity more similar to that of chymotrypsin than trypsin. We report the use of two strategies, substrate phage display and iterative optimization of natural cleavage sites, to identify labile sequences for PSA cleavage. RESULTS: Iterative optimization and substrate phage display converged on the amino-acid sequence SS(Y/F)Y decreases S(G/S) as preferred subsite occupancy for PSA. These sequences were cleaved by PSA with catalytic efficiencies as high as 2200-3100 M-1 s-1, compared with values of 2-46 M-1 s-1 for peptides containing likely physiological target sequences of PSA from the protein semenogelin. Substrate residues that bind to secondary (non-S1) subsites have a critical role in defining labile substrates and can even cause otherwise disfavored amino acids to bind in the primary specificity (S1) pocket. CONCLUSION: The importance of secondary subsites in defining both the specificity and efficiency of cleavage suggests that substrate recognition by PSA is mediated by an extended binding site. Elucidation of preferred subsite occupancy allowed refinement of the structural model of PSA and should facilitate the development of more sensitive activity-based assays and the design of potent inhibitors.  相似文献   

16.
Pseudomonas carboxyl proteinase (PCP), isolated from Pseudomonas sp. 101, and Xanthomonas carboxyl proteinase (XCP), isolated from Xanthomonas sp. T-22, are the first and second examples of unique carboxyl proteinases [EC 3.4.23.33] which are insensitive to aspartic proteinase inhibitors, such as pepstatin, diazoacetyl-DL-norleucine methylester, and 1,2-epoxy-3(p-nitrophenoxy)propane. The substrate specificities of PCP and XCP were studied using a series of synthetic chromogenic peptide substrates with the general structure, P5-P4-P3-P2-Phe-Nph-P2'-P3' (P5, P4, P3, P2, P2', P3': a variety of amino acids, Nph is p-nitro-L-phenylalanine, and the Phe-Nph bond is cleaved). PCP and XCP were shown to hydrolyze a synthetic substrate, Lys-Pro-Ala-Leu-Phe-Nph-Arg-Leu, most effectively among 28 substrates. The kinetic parameters of this peptide for PCP were Km = 6.3 microM, Kcat = 51.4 s-1, and kcat/Km = 8.16 microM-1.s-1. The kinetic parameters for XCP were Km = 3.6 microM, kcat = 52.2 s-1, and kcat/Km = 14.5 microM-1.s-1. PCP showed a stricter substrate specificity than XCP. That is, the specificity constant (kcat/Km) of each substrate for PCP was in general < 0.5 microM-1.s-1, but was drastically improved by the replacement of Lys by Leu at the P2 position. On the other hand, XCP showed a less stringent substrate specificity, with most of the peptides exhibiting reasonable kcat/Km values (> 1.0 microM-1.s-1). Thus it was found that the substrate specificities of PCP and XCP differ considerably, in spite of the high similarity in their primary structures. In addition, tyrostatin was found to be a competitive inhibitor for XCP, with a Ki value of 2.1 nM, as well as for PCP (Ki = 2.6 nM).  相似文献   

17.
Random mutagenesis and screening for enzymatic activity has been used to engineer horse heart myoglobin to enhance its intrinsic peroxidase activity. A chemically synthesized gene encoding horse heart myoglobin was subjected to successive cycles of PCR random mutagenesis. The mutated myoglobin gene was expressed in Escherichia coli LE392, and the variants were screened for peroxidase activity with a plate assay. Four cycles of mutagenesis and screening produced a series of single, double, triple, and quadruple variants with enhanced peroxidase activity. Steady-state kinetics analysis demonstrated that the quadruple variant T39I/K45D/F46L/I107F exhibits peroxidase activity significantly greater than that of the wild-type protein with k1 (for H2O2 oxidation of metmyoglobin) of 1. 34 x 10(4) M-1 s-1 ( approximately 25-fold that of wild-type myoglobin) and k3 [for reducing the substrate (2, 2'-azino-di-(3-ethyl)benzthiazoline-6-sulfonic acid] of 1.4 x 10(6) M-1 s-1 (1.6-fold that of wild-type myoglobin). Thermal stability of these variants as measured with circular dichroism spectroscopy demonstrated that the Tm of the quadruple variant is decreased only slightly compared with wild-type (74.1 degreesC vs. 76.5 degreesC). The rate constants for binding of dioxygen exhibited by the quadruple variant are identical to the those observed for wild-type myoglobin (kon, 22.2 x 10(-6) M-1 s-1 vs. 22.3 x 10(-6) M-1 s-1; koff, 24.3 s-1 vs. 24.2 s-1; KO2, 0.91 x 10(-6) M-1 vs. 0.92 x 10(-6) M-1). The affinity of the quadruple variant for CO is increased slightly (kon, 0.90 x 10(-6) M-1s-1 vs. 0.51 x 10(-6) M-1s-1; koff, 5.08 s-1 vs. 3.51 s-1; KCO, 1.77 x 10(-7) M-1 vs. 1.45 x 10(-7) M-1). All four substitutions are in the heme pocket and within 5 A of the heme group.  相似文献   

18.
The effect of free fatty acids (FFA) and non-enzymatic glycation on the binding kinetics of dansylsarcosine (DS) to human serum albumin (HSA) was studied using the stopped-flow technique. The influence of FFA on the binding parameters of 25% glycated HSA depended on the type of fatty acid. The addition of stearic, oleic and linoleic acids in a concentration of 0.3 mmol/l showed no inhibitory effects on the association rate constant (k2) value for DS binding to 25% glycated HSA (k2 without FFA: 385 +/- 10 s-1, k2 with FFA > or = 385 +/- 10 s-1). In contrast, shorter chain fatty acids (hexanoic, octanoic, decanoic, lauric and myristic acids) showed marked inhibitory effects for 0.3 mmol/l FFA (k2 range: 233 +/- 32 to 69 +/- 5 s-1) and for 0.6 mmol/l FFA (k2 range: 125 +/- 3 to 20 +/- 4 s-1). The association rate constant (k2) as well as the affinity constant (KA) of DS were markedly affected by glycation: k2 was 686 +/- 61 s-1 for 7% glycated HSA, 385 +/- 10 s-1 for 25% glycated HSA and 209 +/- 12 s-1 for 50% glycated HSA. KA decreased from 6.1 +/- 2.9 x 10(5) M-1 for 7% glycated HSA, to 5.1 +/- 0.1 x 10(5) M-1 for 25% glycated HSA and to 1.3 +/- 0.6 x 10(5) M-1 for 50% glycated HSA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Recent studies noted the contribution of alpha42Tyr to the T-R-dependent UV resonance Raman (UVRR) spectral changes of HbA [Nagai, M., et al. (1996) J. Mol. Struct. 379, 65-75; Huang, S., et al. (1997) Biochemistry 36, 6197-6206], but the observed UVRR changes of the Tyr residue cannot be fully interpreted with alpha42Tyr alone. To identify the remaining contributions, the 235 nm-excited UVRR spectra of Tyr mutant Hbs at alpha140, beta35, and beta145 were investigated here. The Fe-His stretching mode demonstrated that all of these mutant Hbs take the T structure in the deoxy form under these experimental conditions. The UVRR change of the Trp residue of these mutants upon the T-R transition was the same as that in HbA, indicating that the T-R-dependent UVRR change of beta37Trp is not due to stacking with Tyr residues but is due to the formation or destruction of a hydrogen bond. The recombinant Hbs beta35Tyr --> Phe and beta35Tyr --> Thr both exhibited UVRR spectra identical with that of HbA, meaning that beta35Tyr is not responsible. In the spectra of des(beta146His,beta145Tyr)Hb with inositol hexaphosphate, the frequency shift of the Tyr RR bands was the same as that in HbA but the intensity enhancement in the CO form was small, suggesting that beta145Tyr contributes to a part of the intensity change, but scarcely relates to the frequency shift. In the spectra of Hb Rouen (alpha140Tyr --> His), the frequency shifts of bands at 1617 (Y8a) and 1177 (Y9a) cm-1 following ligation were half of those in HbA, while the intensity enhancement was not detected. This result means that alpha140Tyr is responsible for both the frequency shift and the intensity changes. It is suggested that the frequency shift of the Tyr RR bands upon the T --> R transition is due to changes in the hydrogen bonding state of alpha42- and alpha140Tyr and that the intensity enhancement is due to changes in the environment of the penultimate Tyr in both alpha and beta subunits (alpha140 and beta145). These alterations in the vibrational spectra clearly demonstrate which tyrosine residues are involved in the T-R transition as a result of modification of their local environments.  相似文献   

20.
Mutation to the conserved Glu399 or Lys192 caused the rate-limiting step of human liver mitochondrial aldehyde dehydrogenase (ALDH2) to change from deacylation to hydride transfer (Sheikh, S., Ni, L., Hurley, T. D., and Weiner, H. (1997) J. Biol. Chem. 272, 18817-18822). Here we further investigated the role of these two NAD+-ribose-binding residues. The E399Q/K/H/D and K192Q mutants had lower dehydrogenase activity when compared with the native enzyme. No pre-steady state burst of NADH formation was found with the E399Q/K and K192Q enzymes when propionaldehyde was used as the substrate; furthermore, each mutant oxidized chloroacetaldehyde slower than propionaldehyde, and a primary isotope effect was observed for each mutant when [2H]acetaldehyde was used as a substrate. However, no isotope effect was observed for each mutant when alpha-[2H]benzaldehyde was the substrate. A pre-steady state burst of NADH formation was observed for the E399Q/K and K192Q mutants with benzaldehyde, and p-nitrobenzaldehyde was oxidized faster than benzaldehyde. Hence, when aromatic aldehydes were used as substrates, the rate-limiting step remained deacylation for all these mutants. The rate-limiting step remained deacylation for the E399H/D mutants when either aliphatic or aromatic aldehydes were used as substrates. The K192Q mutant displayed a change in substrate specificity, with aromatic aldehydes becoming better substrates than aliphatic aldehydes.  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