首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
In Cu-containing nitrite reductase from Alcaligenes faecalis S-6 the axial methionine ligand of the type-1 site was replaced (M150G) to make the copper ion accessible to external ligands that might affect the enzyme's catalytic activity. The type-1 site optical spectrum of M150G (A(460)/A(600)=0.71) differs significantly from that of the native nitrite reductase (A(460)/A(600)=1.3). The midpoint potential of the type-1 site of nitrite reductase M150G (E(M)=312(+/-5)mV versus hydrogen) is higher than that of the native enzyme (E(M)=213(+/-5)mV). M150G has a lower catalytic activity (k(cat)=133(+/-6)s(-1)) than the wild-type nitrite reductase (k(cat)=416(+/-10)s(-1)). The binding of external ligands to M150G restores spectral properties, midpoint potential (E(M)<225mV), and catalytic activity (k(cat)=374(+/-28)s(-1)). Also the M150H (A(460)/A(600)=7.7, E(M)=104(+/-5)mV, k(cat)=0.099(+/-0.006)s(-1)) and M150T (A(460)/A(600)=0.085, E(M)=340(+/-5)mV, k(cat)=126(+/-2)s(-1)) variants were characterized. Crystal structures show that the ligands act as allosteric effectors by displacing Met62, which moves to bind to the Cu in the position emptied by the M150G mutation. The reconstituted type-1 site has an otherwise unaltered geometry. The observation that removal of an endogenous ligand can introduce allosteric control in a redox enzyme suggests potential for structural and functional flexibility of copper-containing redox sites.  相似文献   

2.
cDNAs of the two-domain arginine kinase (AK) (contiguous dimer; denoted by 2D/WT) and its separated domains 1 and 2 (denoted by D1/WT and D2/WT) from the sea anemone Anthopleura japonicus, were cloned into the plasmid pMAL, and recombinant enzymes were expressed in E. coli as MBP fusion proteins. The kinetic parameters kcat, Ka and Kia, were determined for all three AKs. All three enzymes showed distinct AK activity, and had high affinity for arginine (Ka Arg=0.25-0.48 mM). The catalytic efficiency, calculated by kcat/Ka ArgKia ATP, of the 2D/WT enzyme (182 mM(-2)s(-1), the value for one active 40 kDa domain) was two- to three-times higher than values for either D1/WT or D2/WT (80.2 and 86.4mM(-2)s(-1), respectively), suggesting the presence of domain-domain interactions (cooperativity) in the contiguous dimer. The Kia/Ka values of the three enzymes ranged from 0.88 to 1.32, indicating that there is no strong synergism in substrate binding, as seen in typical AKs. Asp62 and Arg193, which are conserved in most AKs and play a key role in stabilizing the substrate-bound structure, are also conserved in the two domains of Anthopleura AK. We replaced Asp62 in D2/WT with Glu or Gly. The catalytic efficiency and Kia/Ka for the D62E mutant were comparable to those of D2/WT, but catalytic efficiency for the D62G mutant was decreased to 13% of that of the D2/WT with a significantly increased value of Kia/Ka (1.92), indicating that Asp62 plays an important role in the expression of AK activity.  相似文献   

3.
The somatic isoform of angiotensin-converting enzyme (ACE) consists of two homologous domains (N- and C-domains), each bearing a catalytic site. We have used the two-domain ACE form and its individual domains to compare characteristics of different domains and to probe mutual functioning of the two active sites within a bovine ACE molecule. The substrate Cbz-Phe-His-Leu (N-carbobenzoxy-L-phenylalanyl-L-histidyl-L-leucine; from the panel of seven) was hydrolyzed faster by the N-domain, the substrates FA-Phe-Gly-Gly (N-(3-[2-furyl]acryloyl)-L-phenylalanyl-glycyl-glycine) and Hip-His-Leu (N-benzoyl-glycyl-L-histidyl-L-leucine) were hydrolyzed by both domains with equal rates, while other substrates were preferentially hydrolyzed by the C-domain. The inhibitor captopril ((2S)-1-(3-mercapto-2-methylpropionyl)-L-proline) bound to the N-domain more effectively than to the C-domain, whereas lisinopril ((S)-N(alpha)-(1-carboxy-3-phenylpropyl)-L-lysyl-L-proline) bound to equal extent with all ACE forms. However, active site titration with lisinopril assayed by hydrolysis of FA-Phe-Gly-Gly revealed that 1 mol of inhibitor/mol of enzyme abolished the activity of either two-domain or single-domain ACE forms, indicating that a single active site functions in bovine somatic ACE. Neither of the k(cat) values obtained for somatic enzyme was the sum of k(cat) values for individual domains, but in every case the value of the catalytic constant of the hydrolysis of the substrate by the two-domain ACE represented the mean quantity of the values of the corresponding catalytic constants obtained for single-domain forms. The results indicate that the two active sites within bovine somatic ACE exhibit strong negative cooperativity.  相似文献   

4.
In the N-terminal domain of thermolysin, two anti-parallel β-strands, Asn112-Ala113-Phe114-Trp115 and Ser118-Gln119-Met120-Val121-Tyr122 are connected by an Asn116-Gly117 turn to form a β-hairpin structure. In this study, we examined the role of Asn116 in the activity and stability of thermolysin by site-directed mutagenesis. Of the 19 Asn116 variants, four (N116A, N116D, N116T and N116Q) were produced in Escherichia coli, by co-expressing the mature and pro domains separately, while the other 15 were not. In the hydrolysis of N-[3-(2-furyl)acryloyl]-glycyl-l-leucine amide (FAGLA) at 25°C, the intrinsic k(cat)/K(m) value of N116D was 320% of that of the wild-type thermolysin (WT), and in the hydrolysis of N-carbobenzoxy-l-aspartyl-l-phenylalanine methyl ester (ZDFM) at pH 7.5 at 25°C, the k(cat)/K(m) value of N116D was 140% of that of WT, indicating that N116D exhibited higher activity than WT. N116Q exhibited similar activity as WT, and N116A and N116T exhibited reduced activities. The first-order rate constants, k(obs), of the thermal inactivation at 80°C were in the order N116A, N116D, N116T > N116Q > WT at all CaCl(2) concentrations examined (1-100 mM), indicating that all variants exhibited reduced stabilities. These results suggest that Asn116 plays an important role in the activity and stability of thermolysin presumably by stabilizing this β-hairpin structure.  相似文献   

5.
The joint substitution of three active-site residues in Escherichia coli (L)-aspartate aminotransferase increases the ratio of l-cysteine sulfinate desulfinase to transaminase activity 10(5)-fold. This change in reaction specificity results from combining a tyrosine-shift double mutation (Y214Q/R280Y) with a non-conservative substitution of a substrate-binding residue (I33Q). Tyr214 hydrogen bonds with O3 of the cofactor and is close to Arg374 which binds the α-carboxylate group of the substrate; Arg280 interacts with the distal carboxylate group of the substrate; and Ile33 is part of the hydrophobic patch near the entrance to the active site, presumably participating in the domain closure essential for the transamination reaction. In the triple-mutant enzyme, k(cat)' for desulfination of l-cysteine sulfinate increased to 0.5s(-1) (from 0.05s(-1) in wild-type enzyme), whereas k(cat)' for transamination of the same substrate was reduced from 510s(-1) to 0.05s(-1). Similarly, k(cat)' for β-decarboxylation of l-aspartate increased from<0.0001s(-1) to 0.07s(-1), whereas k(cat)' for transamination was reduced from 530s(-1) to 0.13s(-1). l-Aspartate aminotransferase had thus been converted into an l-cysteine sulfinate desulfinase that catalyzes transamination and l-aspartate β-decarboxylation as side reactions. The X-ray structures of the engineered l-cysteine sulfinate desulfinase in its pyridoxal-5'-phosphate and pyridoxamine-5'-phosphate form or liganded with a covalent coenzyme-substrate adduct identified the subtle structural changes that suffice for generating desulfinase activity and concomitantly abolishing transaminase activity toward dicarboxylic amino acids. Apparently, the triple mutation impairs the domain closure thus favoring reprotonation of alternative acceptor sites in coenzyme-substrate intermediates by bulk water.  相似文献   

6.
Suzuki T  Tomoyuki T  Uda K 《FEBS letters》2003,533(1-3):95-98
Arginine kinase (AK) from the clam Corbicula japonica is a unique enzyme in that it has an unusual two-domain structure with molecular mass of 80 kDa. It lacks two functionally important amino acid residues, Asp-62 and Arg-193, which are conserved in other 40 kDa AKs and are assumed to be key residues for stabilizing the substrate-bound structure. K m arg and Vmax values for the recombinant two-domain AK were determined. These values were close to those of usual 40 kDa AKs, although Corbicula AK lacks the functionally important Asp-62 and Arg-193. Domain 2 of Corbicula AK was separated from the two-domain enzyme and was expressed in Escherichia coli. Domain 2 still exhibited activity. However, kinetic parameters for domain 2 appeared to be slightly, but significantly, different from those of two-domain AK. Thus, it is likely that the formation of the contiguous dimer alters the kinetic properties of its constituent domains significantly. Comparison of K d arg and K m arg for two-domain AK and its domain 2 showed that the affinity of the enzyme for arginine is greater in the presence of substrate ATP than in its absence. Presumably this difference is correlated with the large structural differences in the enzyme in the presence or absence of substrate, namely open and closed structures. We expressed three mutants of Corbicula AK domain 2 (His-60 to Gly or Arg, Asp-197 to Gly), and determined their K m arg and Vmax values. The affinity for the substrate arginine in mutant enzymes was reduced considerably, accompanied by a decrease in Vmax. These results suggest that His-60 and Asp-197 affect the substrate binding system, and are consistent with the hypothesis that a hydrogen bond is formed between His-60 and Asp-197 in Corbicula AK as a substitute for the Asp-62 and Arg-193 bond in normal AKs.  相似文献   

7.
Wang GP  Hansen MR  Grubmeyer C 《Biochemistry》2012,51(22):4406-4415
Residue-to-alanine mutations and a two-amino acid deletion have been made in the highly conserved catalytic loop (residues 100-109) of Salmonella typhimurium OMP synthase (orotate phosphoribosyltransferase, EC 2.4.2.10). As described previously, the K103A mutant enzyme exhibited a 10(4)-fold decrease in k(cat)/K(M) for PRPP; the K100A enzyme suffered a 50-fold decrease. Alanine mutations at His105 and Glu107 produced 40- and 7-fold decreases in k(cat)/K(M), respectively, and E101A, D104A, and G106A were slightly faster than the wild-type (WT) in terms of k(cat), with minor effects on k(cat)/K(M). Equilibrium binding of OMP or PRPP in binary complexes was affected little by loop mutation, suggesting that the energetics of ground-state binding have little contribution from the catalytic loop, or that a favorable binding energy is offset by costs of loop reorganization. Pre-steady-state kinetics for mutants showed that K103A and E107A had lost the burst of product formation in each direction that indicated rapid on-enzyme chemistry for WT, but that the burst was retained by H105A. Δ102Δ106, a loop-shortened enzyme with Ala102 and Gly106 deleted, showed a 10(4)-fold reduction of k(cat) but almost unaltered K(D) values for all four substrate molecules. The 20% (i.e., 1.20) intrinsic [1'-(3)H]OMP kinetic isotope effect (KIE) for WT is masked because of high forward and reverse commitment factors. K103A failed to express intrinsic KIEs fully (1.095 ± 0.013). In contrast, H105A, which has a smaller catalytic lesion, gave a [1'-(3)H]OMP KIE of 1.21 ± 0.0005, and E107A (1.179 ± 0.0049) also gave high values. These results are interpreted in the context of the X-ray structure of the complete substrate complex for the enzyme [Grubmeyer, C., Hansen, M. R., Fedorov, A. A., and Almo, S. C. (2012) Biochemistry 51 (preceding paper in this issue, DOI 10.1021/bi300083p )]. The full expression of KIEs by H105A and E107A may result from a less secure closure of the catalytic loop. The lower level of expression of the KIE by K103A suggests that in these mutant proteins the major barrier to catalysis is successful closure of the catalytic loop, which when closed, produces rapid and reversible catalysis.  相似文献   

8.
Histamine dehydrogenase (NSHADH) can be isolated from cultures of Nocardioides simplex grown with histamine as the sole nitrogen source. A previous report suggested that NSHADH might contain the quinone cofactor tryptophan tryptophyl quinone (TTQ). Here, the hdh gene encoding NSHADH is cloned from the genomic DNA of N. simplex, and the isolated enzyme is subjected to a full spectroscopic characterization. Protein sequence alignment shows NSHADH to be related to trimethylamine dehydrogenase (TMADH: EC 1.5.99.7), where the latter contains a bacterial ferredoxin-type [4Fe-4S] cluster and 6-S-cysteinyl FMN cofactor. NSHADH has no sequence similarity to any TTQ containing amine dehydrogenases. NSHADH contains 3.6+/-0.3 mol Fe and 3.7+/-0.2 mol acid labile S per subunit. A comparison of the UV/vis spectra of NSHADH and TMADH shows significant similarity. The EPR spectrum of histamine reduced NSHADH also supports the presence of the flavin and [4Fe-4S] cofactors. Importantly, we show that NSHADH has a narrow substrate specificity, oxidizing only histamine (K(m)=31+/-11 microM, k(cat)/K(m)=2.1 (+/-0.4)x10(5)M(-1)s(-1)), agmatine (K(m)=37+/-6 microM, k(cat)/K(m)=6.0 (+/-0.6)x10(4)M(-1)s(-1)), and putrescine (K(m)=1280+/-240 microM, k(cat)/K(m)=1500+/-200 M(-1)s(-1)). A kinetic characterization of the oxidative deamination of histamine by NSHADH is presented that includes the pH dependence of k(cat)/K(m) (histamine) and the measurement of a substrate deuterium isotope effect, (D)(k(cat)/K(m) (histamine))=7.0+/-1.8 at pH 8.5. k(cat) is also pH dependent and has a reduced substrate deuterium isotope of (D)(k(cat))=1.3+/-0.2.  相似文献   

9.
Stratton JR  Pelton JG  Kirsch JF 《Biochemistry》2001,40(35):10411-10416
The low-barrier hydrogen bond (LBHB) between the Asp and His residues of the catalytic triad in a serine protease was perturbed via the D32C mutation in subtilisin BPN' (Bacillus protease N'). This mutant enzyme catalyzes the hydrolysis of N-Suc-Ala-Ala-Pro-Phe-SBzl with a k(cat)/K(m) value that is only 8-fold reduced from that of the wild-type (WT) enzyme. The value of k(cat)/K(m) for the corresponding p-nitroanilide (pNA) substrate is only 50-fold lower than that of the WT enzyme (DeltaDeltaG++ = 2.2 kcal/mol). The pK(a) controlling the ascending limb of the pH versus k(cat)/K(m) profile is lowered from 7.01 (WT) to 6.53 (D32C), implying that any hydrogen bond replacing that between Asp32 and His64 of the WT enzyme most likely involves the neutral thiol rather than the thiolate form of Cys32. It is shown by viscosity variation that the reaction of WT subtilisin with N-Suc-Ala-Ala-Pro-Phe-SBzl is 50% (sucrose) to 100% (glycerol) diffusion-controlled, while that of the D32C construct is 29% (sucrose) to 76% (glycerol) diffusion-controlled. The low-field NMR resonance of 18 ppm that has been assigned to a proton shared by Asp32 and His64, and is considered diagnostic of a LBHB in the WT enzyme, is not present in D32C subtilisin. Thus, the LBHB is not an inherent requirement for substantial rate enhancement for subtilisin.  相似文献   

10.
Mueller M  Nidetzky B 《FEBS letters》2007,581(7):1403-1408
Replacements of Asp-295 by Asn (D295N) and Glu (D295E) decreased the catalytic center activity of Leuconostoc mesenteroides sucrose phosphorylase to about 0.01% of the wild-type level (k(cat)=200s(-1)). Glucosylation and deglucosylation steps of D295N were affected uniformly, approximately 10(4.3)-fold, and independently of leaving group ability and nucleophilic reactivity of the substrate, respectively. pH dependences of the catalytic steps were similar for D295N and wild-type. The 10(5)-fold preference of the wild-type for glucosyl transfer compared with mannosyl transfer from phosphate to fructose was lost in D295N and D295E. Selective disruption of catalysis to glucosyl but not mannosyl transfer in the two mutants suggests that the side chain of Asp-295, through a strong hydrogen bond with the equatorial sugar 2-hydroxyl, stabilizes the transition states flanking the beta-glucosyl enzyme intermediate by > or = 23kJ/mol.  相似文献   

11.
The function of the N-terminal domain ( approximately 350 residues) of the Pta (phosphotransacetylase) enzyme of Salmonella enterica is unclear. Results from in vivo genetic and in vitro studies suggest that the N-terminal domain of Pta is a sensor for NADH and pyruvate. We isolated 10 single-amino acid variants of Pta that, unlike the wild-type protein, supported growth of a strain of S. enterica devoid of Acs (acetyl-CoA synthetase; AMP-forming) activity on 10 mm acetate. All mutations were mapped within the N-terminal domain of the protein. Kinetic analyses of the wild type and three variant Pta proteins showed that two of the variant proteins were faster enzymes (k(cat) 2.5-3-fold > k(cat) Pta(WT). Results from sedimentation equilibrium experiments are consistent with Pta(WT) being a trimer. Pta variants formed more hexamer than the Pta(WT) protein. NADH inhibited Pta(WT) activity by inducing a conformational change detectable by limited trypsin proteolysis; NADH did not inhibit variant protein Pta(R252H). Pyruvate stimulated Pta(WT) activity, and its effect was potentiated in the variants, being most pronounced on Pta(R252H).  相似文献   

12.
An intramolecularly quenched fluorogenic peptide containing o-aminobenzoyl (Abz) and ethylenediamine 2,4-dinitrophenyl (Eddnp) groups at amino- and carboxyl-terminal amino acid residues, Abz-Lys-Pro-Ile-Glu-Phe-Phe-Arg-Leu-Eddnp, was hydrolyzed by purified human pepsin, gastricsin, and gastric juice uniquely at the Phe-Phe bond. Kinetic parameters determined for purified pepsin were K(m)=0.68+/-0.11 microM; k(cat)=6.3+/-0.16s(-1); k(cat)/K(m)=9.26s(-1) microM(-1); Gastricsin showed K(m)=2.69+/-0.18 microM; k(cat)=0.03+/-0.005s(-1); k(cat)/K(m)=0.011s(-1) microM(-1). Gastric juice (21 samples) from subjects without gastric disorders at endoscopy examination showed activities varying from 0.0008 to 9.72 micromolml(-1)min(-1). Pepstatin A inhibition of gastric juice enzymatic activity was complete at 3.4x10(-5)M (final concentration) inhibitor. In the proposed method the presence of a unique scissile bond in the synthetic substrate provides a direct ratio between enzymatic activity and amount of substrate hydrolyzed, and a unique step reaction facilitates the use of this assay for the determination of the activity of aspartic proteinases in biological fluids and during enzyme purification procedures.  相似文献   

13.
A recombinant form of BACE1 (β-site amyloid precursor protein cleaving enzyme-1) corresponding to positions 46-454 of the extracellular domain of the original membrane enzyme was prepared. The recombinant BACE1 (rBACE1) had the kinetic parameters K(m)=5.5μM and k(cat)=1719s(-1). Using several libraries of substrates containing unnatural amino acids, the specificity of rBACE1 was evaluated. LC/MS of digests derived from the libraries clarified that a dodecapeptide containing unnatural amino acids at P(2) to [Formula: see text] was a superior cleavage sequence.  相似文献   

14.
Pre-steady state, stopped flow analysis of Escherichia coli D-3-phosphoglycerate dehydrogenase was performed by following the fluorescence of protein tryptophan and the fluorescence resonance energy transfer from protein tryptophan to bound NADH. The results indicate that binding of substrates is ordered, with coenzyme, NADH, binding first. Furthermore, the analysis indicated that there are two sets of sites on the tetrameric enzyme that can be differentiated by their kinetic behavior. NADH binding was consistent with an initial binding event followed by a slow conformational change for each site. The slow conformational change is responsible for the apparent tight binding of NADH to the apoenzyme but is too slow to participate in the catalytic cycle when the enzyme is rapidly turning over. Subsequent binding of the substrate, alpha-ketoglutarate, was characterized by a rapid equilibrium binding event followed by a conformational change for each site. Catalysis in the direction of NAD(+) reduction showed a distinct burst of activity followed by a slow rate of turnover, indicating that the rate-limiting step is after hydride transfer. Catalysis in the direction of NADH oxidation did not display burst kinetics, indicating that the rate-limiting step is at or before the hydride transfer step. The burst data indicated that the rate of NAD(+) reduction (3.8 s(-1)) is similar to the k(cat) of the enzyme (2-3 s(-1)) in that direction. However, analysis of the reaction with deuterated NADH failed to show an effect on the velocity of the reaction with a V(H)/V(D)=1.07+/-0.06. None of the other rates determined by stopped flow analysis could account for the k(cat) of the enzyme in either direction (forward k(cat)=0.01 s(-1), reverse k(cat)=2-3 s(-1)), suggesting that the rate-limiting step in both directions is a conformational change in the enzyme that is not detected optically.  相似文献   

15.
The Plasmodium falciparum cysteine protease falcipain-2 is a trophozoite hemoglobinase and potential antimalarial drug target. Unlike other studied papain family proteases, falcipain-2 does not require its prodomain for folding to active enzyme. Rather, folding is mediated by an amino-terminal extension of the mature protease. As in related enzymes, the prodomain is a potent inhibitor of falcipain-2. We now report further functional evaluation of the domains of falcipain-2 and related plasmodial proteases. The minimum requirement for folding of falcipain-2 and four related plasmodial cysteine proteases was inclusion of a 14-15-residue amino-terminal folding domain, beginning with a conserved Tyr. Chimeras of the falcipain-2 catalytic domain with extensions from six other plasmodial proteases folded normally and had kinetic parameters (k(cat)/K(m) 124,000-195,000 M(-1) s(-1)) similar to those of recombinant falcipain-2 (k(cat)/K(m) 120,000 M(-1) s(-1)), indicating that the folding domain is functionally conserved across the falcipain-2 subfamily. Correct folding also occurred when the catalytic domain was refolded with a separate prodomain-folding domain construct but not with an isolated folding domain peptide. Thus, the prodomain mediated interaction between the other two domains when they were not covalently bound. The prodomain-catalytic domain interaction was independent of the active site, because it was blocked by free inactive catalytic domain but not by the active site-binding peptide leupeptin. The folded catalytic domain retained activity after purification from the prodomain-folding domain construct (k(cat)/K(m) 168,000 M(-1) s(-1)), indicating that the folding domain is not required for activity once folding has been achieved. Activity was lost after nonreducing gelatin SDS-PAGE but not native gelatin PAGE, indicating that correct disulfide bonds are insufficient to direct appropriate folding. Our results identify unique features of the falcipain-2 subfamily with independent mediation of activity, folding, and inhibition.  相似文献   

16.
Genomic analysis of a hyperthermophilic archaeon, Thermococcus sp. NA1, revealed the presence of an 1,497 bp open reading frame, encoding a protein of 499 amino acids. The deduced amino acid sequence was similar to thermostable carboxypeptidase 1 from Pyrococcus furiosus, a member of peptidase family M32. Five motifs, including the HEXXH motif with two histidines coordinated with the active site metal, were conserved. The carboxypeptidase gene was cloned and overexpressed in Escherichia coli. Molecular masses assessed by SDS-PAGE and gel filtration were 61 kDa and 125 kDa respectively, which points to a dimeric structure for the recombinant enzyme, designated TNA1_CP. The enzyme showed optimum activity toward Z-Ala-Arg at pH 6.5 and 70-80 degrees C (k(cat)/K(m)=8.3 mM(-1) s(-1)). In comparison with that of P. furiosus CP (k(cat)/K(m)=667 mM(-1) s(-1)), TNA1_CP exhibited 80-fold lower catalytic efficiency. The enzyme showed broad substrate specificity with a preference for basic, aliphatic, and aromatic C-terminal amino acids. This broad specificity was confirmed by C-terminal ladder sequencing of porcine N-acetyl-renin substrate by TNA1_CP.  相似文献   

17.
The pikromycin biosynthetic gene cluster contains the pikAV gene encoding a type II thioesterase (TEII). TEII is not responsible for polyketide termination and cyclization, and its biosynthetic role has been unclear. During polyketide biosynthesis, extender units such as methylmalonyl acyl carrier protein (ACP) may prematurely decarboxylate to generate the corresponding acyl-ACP, which cannot be used as a substrate in the condensing reaction by the corresponding ketosynthase domain, rendering the polyketide synthase module inactive. It has been proposed that TEII may serve as an "editing" enzyme and reactivate these modules by removing acyl moieties attached to ACP domains. Using a purified recombinant TEII we have tested this hypothesis by using in vitro enzyme assays and a range of acyl-ACP, malonyl-ACP, and methylmalonyl-ACP substrates derived from either PikAIII or the loading didomain of DEBS1 (6-deoxyerythronolide B synthase; AT(L)-ACP(L)). The pikromycin TEII exhibited high K(m) values (>100 microm) with all substrates and no apparent ACP specificity, catalyzing cleavage of methylmalonyl-ACP from both AT(L)-ACP(L) (k(cat)/K(m) 3.3 +/- 1.1 m(-1) s(-1)) and PikAIII (k(cat)/K(m) 2.9 +/- 0.9 m(-1) s(-1)). The TEII exhibited some acyl-group specificity, catalyzing hydrolysis of propionyl (k(cat)/K(m) 15.8 +/- 1.8 m(-1) s(-1)) and butyryl (k(cat)/K(m) 17.5 +/- 2.1 m(-1) s(-1)) derivatives of AT(L)-ACP(L) faster than acetyl (k(cat)/K(m) 4.9 +/- 0.7 m(-1) s(-1)), malonyl (k(cat)/K(m) 3.9 +/- 0.5 m(-1) s(-1)), or methylmalonyl derivatives. PikAIV containing a TEI domain catalyzed cleavage of propionyl derivative of AT(L)-ACP(L) at a dramatically lower rate than TEII. These results provide the first unequivocal in vitro evidence that TEII can hydrolyze acyl-ACP thioesters and a model for the action of TEII in which the enzyme remains primarily dissociated from the polyketide synthase, preferentially removing aberrant acyl-ACP species with long half-lives. The lack of rigorous substrate specificity for TEII may explain the surprising observation that high level expression of the protein in Streptomyces venezuelae leads to significant (>50%) titer decreases.  相似文献   

18.
A kinetic study of CO(2) hydration was carried out using the water-soluble zinc model complex with water-soluble nitrilotris(2-benzimidazolylmethyl-6-sulfonate) L1S, [L1SZn(OH(2))](-), mimicking the active site of carbonic anhydrase, in the presence and absence of anion inhibitors NCS(-) and Cl(-). The obtained rate constants k(cat) for CO(2) hydration were 5.9x10(2), 1. 7x10(3), and 3.1x10(3) M(-1) s(-1) at 5, 10, and 15 degrees C, respectively: the k(cat)=ca. 10(4) M(-1) s(-1) extrapolated towards 25 degrees C has been the largest among the reported k(cat) using zinc model complexes for carbonic anhydrase. It was also revealed that NCS(-), Cl(-) and acetazolamide play a role of inhibitors by the decrease of k(cat): 7x10(2) and 2x10(3) M(-1) s(-1) for NCS(-) and Cl(-) at 15 degrees C, respectively. The sequence of their magnitudes in k(cat) is Cl(-) approximately acetazolamide>NCS(-), where the sequence Cl(-)>NCS(-) is confirmed for native carbonic anhydrase. The difference of k(cat) or k(obs) between NCS(-) and Cl(-) resulted from that between the stability constants K(st)=2x10(3) for [L1SZn(NCS)](2-) and 1x10(2) M(-1) for [L1SZnCl](2-) in D(2)O: for water-insoluble tris(2-benzimidazolylmethyl)amine L1, K(st)=1.8x10(4) for [L1Zn(NCS)](2-) and 1.5x10(3) M(-1) for [L1ZnCl](2-)in CD(3)CN/D(2)O (50% v/v). The crystal structure of anion-binding zinc model complexes [L1Zn(OH(2))](0.5)[L1ZnCl](0.5) (ClO(4))(1.5) 1(0.5)2(0.5)(ClO(4))(1.5) was revealed by X-ray crystallography. The geometry around Zn(2+) in 1 and 2 was tetrahedrally coordinated by three benzimidazolyl nitrogen atoms and one oxygen atom of H(2)O, or Cl(-).  相似文献   

19.
Thermolysin is remarkably activated and stabilized by neutral salts, and surface charges are suggested important in its activity and stability. The effects of introducing negative charge into the molecular surface on its activity and stability are described. Seven serine residues were selected, and each of them was changed for aspartate by site-directed mutagenesis in a thermolysin mutant. In the hydrolysis of N-[3-(2-furyl)acryloyl]-glycyl-l-leucine amide, the k(cat)/K(m) values of all mutants were almost similar to that of the wild-type enzyme (WT). However, those of six out of seven mutants were enhanced 17-19 times with 4 M NaCl, being slightly higher than WT. The remaining casein-hydrolyzing activities of the S53D and S65D mutants (Ser53 and Ser65 are replaced with Asp, respectively) after 30-min incubation with 10 mM CaCl(2) at 85 degrees C were 78 and 63%, being higher than those of WT (51%) and the other mutants (35-53%). S53D was stabilized with increase in the enthalpy change of activation for thermal inactivation while S65D was with decrease in the entropy change of activation. The stability of WT was enhanced by CaCl(2) and reached the level of S53D and S65D at 100 mM, suggesting that S53D and S65D might be stabilized by reinforcement of the Ca(2+)-binding structures.  相似文献   

20.
Inverse thinking about double mutants of enzymes   总被引:3,自引:0,他引:3  
Mildvan AS 《Biochemistry》2004,43(46):14517-14520
The quantitative effect of a second damaging mutation on a mutated enzyme may be additive, partially additive, synergistic, antagonistic, or absent, in the double mutant. Each of these five possible types of interactions has its own mechanistic explanation [Mildvan, A. S., Weber, D. J., and Kuliopulos, A. (1992) Arch. Biochem. Biophys. 294, 327-340]. Additive effects indicate independent functioning of the two residues in the process being studied, such as catalysis (k(cat)) or substrate binding (K(S)). Departures from additivity reflect interaction of the two residues. Thus, partial additivity indicates cooperativity, synergy indicates anticooperativity, and antagonism indicates opposing structural effects of the two mutations. No additional effects represent limiting cases of either partial additivity or antagonism. A significant conceptual simplification is achieved by applying inverse thinking, namely, by using the parameters of the double mutant rather than those of the wild-type enzyme as the reference point. To explain partially additive effects on k(cat), inverse thinking starts with the k(cat) of the double mutant. Restoring only one residue increases k(cat) by the factor A. Restoring only the other residue increases k(cat) by the factor B. Restoring both residues is shown to increase k(cat) by a factor greater than A x B, with the excess directly measuring the cooperativity. Similarly, inverse thinking provides simpler and more intuitive explanations of synergistic and antagonistic effects, as illustrated by specific examples.  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