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 265 毫秒
1.
Laccase from Myceliophthora thermophila (MtL) was expressed in functional form in Saccharomyces cerevisiae. Directed evolution improved expression eightfold to the highest yet reported for a laccase in yeast (18 mg/liter). Together with a 22-fold increase in k(cat), the total activity was enhanced 170-fold. Specific activities of MtL mutants toward 2,2'-azino-bis(3-ethylbenzthiazoline-6-sulfonic acid) and syringaldazine indicate that substrate specificity was not changed by the introduced mutations. The most effective mutation (10-fold increase in total activity) introduced a Kex2 protease recognition site at the C-terminal processing site of the protein, adjusting the protein sequence to the different protease specificities of the heterologous host. The C terminus is shown to be important for laccase activity, since removing it by a truncation of the gene reduces activity sixfold. Mutations accumulated during nine generations of evolution for higher activity decreased enzyme stability. Screening for improved stability in one generation produced a mutant more stable than the heterologous wild type and retaining the improved activity. The molecular mass of MtL expressed in S. cerevisiae is 30% higher than that of the same enzyme expressed in M. thermophila (110 kDa versus 85 kDa). Hyperglycosylation, corresponding to a 120-monomer glycan on one N-glycosylation site, is responsible for this increase. This S. cerevisiae expression system makes MtL available for functional tailoring by directed evolution.  相似文献   

2.
Nguyen KT  Pei D 《Biochemistry》2005,44(23):8514-8522
N-Formyl peptides are derived from proteolytic degradation/processing of bacterial and mitochondrial proteins and serve as potent chemoattractants for mammalian phagocytic leukocytes. A response to the chemotactic N-formyl peptides released by commensal bacteria in the gut region could be detrimental, leading to unwanted inflammation. Here, two enzymes that act sequentially to degrade N-formyl peptides were purified from the rat intestinal mucosal layer and biochemically characterized. The first enzyme cleaves chemotactic peptide f-MLF to release N-formylmethionine (f-Met) and dipeptide leucylphenylalanine, with a k(cat) value of 14 s(-)(1), a K(M) value of 0.60 mM, and a k(cat)/K(M) value of 22 500 M(-)(1) s(-)(1). In-gel tryptic digestion followed by mass spectral fingerprinting identified the protein as the alpha-N-acylpeptide hydrolase (or acylamino acid-releasing enzyme, EC 3.4.19.1). The second enzyme hydrolyzes N-formylmethionine into formate and methionine with a k(cat) value of 7.9 s(-)(1), a K(M) value of 3.1 mM, and a k(cat)/K(M) value of 2550 M(-)(1) s(-)(1). This protein was identified as the N-acylase IA (or N(alpha)-acyl-l-amino acid amidohydrolase, EC 3.5.1.14). Together, these two enzymes play a protective role in degrading bacterial and mitochondrial N-formylated peptides.  相似文献   

3.
There are several known routes for the metabolic detoxication of alpha,beta-unsaturated aldehydes and ketones, including conjugation to glutathione and reduction and oxidation of the aldehyde to an alcohol and a carboxylic acid, respectively. In this study, we describe a fourth class of detoxication that involves the reduction of the alpha,beta-carbon=carbon double bond to a single bond. This reaction is catalyzed by NAD(P)H-dependent alkenal/one oxidoreductase (AO), an enzyme heretofore known as leukotriene B4 12-hydroxydehydrogenase, 15-oxoprostaglandin 13-reductase, and dithiolethione-inducible gene-1. AO is shown to effectively reduce cytotoxic lipid peroxidation products such as 4-hydroxy-2-nonenal (HNE) (k(cat) = 4.0 x 10(3) min(-1); k(cat)/K(m) = 3.3 x 10(7) min(-1) M(-1)) and acrolein (k(cat) = 2.2 x 10(2) min(-1); k(cat)/K(m) = 1.5 x 10(6) min(-1) M(-1)) and common industrial compounds such as ethyl vinyl ketone (k(cat) = 9.6 x 10(3) min(-1); k(cat)/K(m) = 8.8 x 10(7) min(-1) M(-1)) and 15-oxoprostaglandin E1 (k(cat) = 2.4 x 10(3) min(-1); k(cat)/K(m) = 2.4 x 10(9) min(-1) M(-1)). Furthermore, transfection of human embryonic kidney cells with a rat liver AO expression vector protected these cells from challenge with HNE. The concentration of HNE at which 50% of the cells were killed after 24 h increased from approximately 15 microM in control cells to approximately 70 microM in AO-transfected cells. Overexpression of AO also completely abolished protein alkylation by HNE at all concentrations tested (up to 30 microM). Thus, we describe a novel antioxidative activity of a previously characterized bioactive lipid-metabolizing enzyme that could prove to be therapeutically or prophylactically useful due to its high catalytic rate and inducibility.  相似文献   

4.
Stewart RC  Jahreis K  Parkinson JS 《Biochemistry》2000,39(43):13157-13165
The histidine protein kinase CheA plays a central role in the bacterial chemotaxis signal transduction pathway. Autophosphorylated CheA passes its phosphoryl group to CheY very rapidly (k(cat) approximately 750 s(-)(1)). Phospho-CheY in turn influences the direction of flagellar rotation. The autophosphorylation site of CheA (His(48)) resides in its N-terminal P1 domain. The adjacent P2 domain provides a high-affinity binding site for CheY, which might facilitate the phosphotransfer reaction by tethering CheY in close proximity to the phosphodonor located in P1. To explore the contribution of P2 to the CheA --> CheY phosphotransfer reaction in the Escherichia coli chemotaxis system, we examined the transfer kinetics of a mutant CheA protein (CheADeltaP2) in which the 98 amino acid P2 domain had been replaced with an 11 amino acid linker. We used rapid-quench and stopped-flow fluorescence experiments to monitor phosphotransfer to CheY from phosphorylated wild-type CheA and from phosphorylated CheADeltaP2. The CheADeltaP2 reaction rates were significantly slower and the K(m) value was markedly higher than the corresponding values for wild-type CheA. These results indicate that binding of CheY to the P2 domain of CheA indeed contributes to the rapid kinetics of phosphotransfer. Although phosphotransfer was slower with CheADeltaP2 (k(cat)/K(m) approximately 1.5 x 10(6) M(-)(1) s(-)(1)) than with wild-type CheA (k(cat)/K(m) approximately 10(8) M(-)(1) s(-)(1)), it was still orders of magnitude faster than the kinetics of CheY phosphorylation by phosphoimidazole and other small molecule phosphodonors (k(cat)/K(m) approximately 5-50 M(-)(1) s(-)(1)). We conclude that the P1 domain of CheA also makes significant contributions to phosphotransfer rates in chemotactic signaling.  相似文献   

5.
Meyer-Almes FJ  Auer M 《Biochemistry》2000,39(43):13261-13268
A new mathematical formalism is deduced which allows for the calculation of the k(cat) over K(M) ratio based on measurements of the enzyme kinetics with substrate concentrations much lower than K(M). The equations are also applied on the action of an inhibitor on enzyme activity yielding the binding constant, K(i), of an inhibitor molecule. For practical evaluation of the new theoretical approach, the enzymatic reaction of CD45 phosphatase was used as a well-characterized model system with known inhibitors for testing the K(i) value determination scheme. The k(cat)/K(M) ratio was calulated to be 4.7 x 10(5) M(-)(1) s(-)(1), the K(i) of the inhibitor molecule PKF52-524 was estimated to be (1-2) x 10(-)(7) M and the association rate of the inhibitor PKF52-524 to CD45 phosphatase was estimated to be 59 M(-)(1) s(-)(1).  相似文献   

6.
High-level expression of human liver monoamine oxidase B in Pichia pastoris   总被引:1,自引:0,他引:1  
The high-level heterologous expression, purification, and characterization of the mitochondrial outer membrane enzyme human liver monoamine oxidase B (MAO B) using the methylotrophic yeast Pichia pastoris expression system are described. A 2-L culture of P. pastoris expresses approximately 1700 U of MAO B activity, with the recombinant enzyme associated tightly with the membrane fraction of the cell lysate. By a modification of the published procedure for purification of bovine liver MAO B [Salach, J. I. (1979) Arch. Biochem. Biophys. 192, 128-137], recombinant human liver MAO B is purified in a 34% yield ( approximately 200 mg from 2 L of cell culture). The isolated enzyme exhibits an M(r) of approximately 60, 000 on SDS-PAGE and 59,474 from electrospray mass spectrometry measurements, which is in good agreement with the mass predicted from the gene sequence and inclusion of the covalent FAD. One mole of covalent FAD per mole of MAO B is present in the purified enzyme and is bound by an 8alpha-S-cysteinyl(397) linkage, as identified by electrospray mass spectrometry of the isolated tryptic/chymotryptic flavin peptide. Recombinant human liver MAO B and bovine liver MAO B are shown to be acetylated at the seryl residues at their respective amino termini. The benzylamine oxidase activity of recombinant MAO B ranges from 3.0 to 3.4 U/mg and steady-state kinetic parameters for this enzyme preparation compare well with those published for the bovine liver enzyme: k(cat) = 600 min(-1), K(m)(benzylamine) = 0.50 mM, and K(m)(O(2)) = 0.33 mM. Kinetic isotope effect parameters using [alpha,alpha-(2)H(2)]benzylamine are also similar to those found for the bovine enzyme. Recombinant MAO B exhibits a (D)k(cat) = 4.7, a (D)[k(cat)/K(m)(benzylamine)] = 4.5, and a (D)[k(cat)/K(m)(O(2))] = 1.0. In contrast to bovine liver MAO B, no evidence was found for the presence of any anionic flavin radical either by UV-vis or by EPR spectroscopy in the resting form of the enzyme. These data demonstrate the successful heterologous expression of a functional, membrane-bound MAO B, which will permit a number of mutagenesis studies as structural and mechanistic probes not previously possible.  相似文献   

7.
Salts inhibit the activity of sweet almond beta-glucosidase. For cations (Cl(-) salts) the effectiveness follows the series: Cu(+2), Fe(+2)>Zn(+2)>Li(+)>Ca(+2)>Mg(+2)>Cs(+)>NH(4)(+)>Rb(+)>K(+)>Na(+) and for anions (Na(+) salts) the series is: I(-)>ClO(4)(-)>(-)SCN>Br(-) approximately NO(3)(-)>Cl(-) approximately (-)OAc>F(-) approximately SO(4)(-2). The activity of the enzyme, like that of most glycohydrolases, depends on a deprotonated carboxylate (nucleophile) and a protonated carboxylic acid for optimal activity. The resulting pH-profile of k(cat)/K(m) for the beta-glucosidase-catalyzed hydrolysis of p-nitrophenyl glucoside is characterized by a width at half height that is strongly sensitive to the nature and concentration of the salt. Most of the inhibition is due to a shift in the enzymic pK(a)s and not to an effect on the pH-independent second-order rate constant, (k(cat)/K(m))(lim). For example, as the NaCl concentration is increased from 0.01 M to 1.0 M the apparent pK(a1)increases (from 3.7 to 4.9) and the apparent pK(a2)decreases (from 7.2 to 5.9). With p-nitrophenyl glucoside, the value of the pH-independent (k(cat)/K(m))(lim) (=9 x 10(4) M(-1) s(-1)) is reduced by less than 4% as the NaCl concentration is increased. There is a similar shift in the pK(a)s when the LiCl concentration is increased to 1.0 M. The results of these salt-induced pK(a) shifts rule out a significant contribution of reverse protonation to the catalytic efficiency of the enzyme. At low salt concentration, the fraction of the catalytically active monoprotonated enzyme in the reverse protonated form (i.e., proton on the group with a pK(a) of 3.7 and dissociated from the group with a pK(a) of 7.2) is very small ( approximately 0.03%). At higher salt concentrations, where the two pK(a)s become closer, the fraction of the monoprotonated enzyme in the reverse protonated form increases over 300-fold. However, there is no increase in the intrinsic reactivity, (k(cat)/K(m))(lim), of the monoprotonated species. For other enzymes which may show such salt-induced pK(a) shifts, this provides a convenient test for the role of reverse protonation.  相似文献   

8.
The phosphoglucomutase gene from a wild type Fusarium oxysporum strain (F3), was homologously expressed, under the control of the constitutive promoter of gpdA of Aspergillus nidulans. The transformant produced elevated levels of phosphoglucomutase activity compared to the wild type, a fact that facilitated the subsequent purification procedure. The enzyme (FoPGM) was purified to homogeneity applying three anion exchange and one gel filtration chromatography steps. The native enzyme revealed a monomeric structure with a molecular mass of 60 kDa, while the isoelectric point was 3.5. FoPGM was active in pH ranged from 6.0 to 8.0, with an optimum using 3-(N-morpholino)propanesulfonic acid buffer at 7.0, while loss of activity was observed when phosphate buffer was used in the above mentioned pH range. The optimal temperature for activity was 45°C but the enzyme became unstable at temperatures above 40°C. FoPGM requires the presence of a divalent cation for its function with maximum activity being obtained with Co(2+). The apparent K(m) for Co(2+) was found to be 10 μM. The enzyme was also active with other divalent metal ions such as Mn(2+), Mg(2+), Ni(2+) and Ca(2+) but to a lesser extent. The following kinetic constants were determined: v(max), 0.74 μmol mg(protein)(-1)min(-1); k(cat), 44.2 min(-1); K(m)(G1P), 0.10mM; K(m)(G1,6 diP), 1.03 μM; k(cat)/K(m)(G1P), 443 mM(-1)min(-1) and k(cat)/K(m)(G1,6 diP), 42,860 mM(-1)min(-1). The enzyme was considered to follow a Ping Pong substituted enzyme or enzyme isomerization mechanism.  相似文献   

9.
Klimacek M  Nidetzky B 《Biochemistry》2002,41(31):10158-10165
Mannitol dehydrogenases (MDH) are a family of Zn(2+)-independent long-chain alcohol dehydrogenases that catalyze the regiospecific NAD(+)-dependent oxidation of a secondary alcohol group in polyol substrates. pH and primary deuterium kinetic isotope effects on kinetic parameters for reaction of recombinant MDH from Pseudomonas fluorescens with D-mannitol have been measured in H(2)O and D(2)O at 25 degrees C and used to determine the relative timing of C-H and O-H bond cleavage steps during alcohol conversion. The enzymatic rates decreased at low pH; apparent pK values for log(k(cat)/K(mannitol)) and log k(cat) were 9.2 and 7.7 in H(2)O, respectively, and both were shifted by +0.4 pH units in D(2)O. Proton inventory plots for k(cat) and k(cat)/K(mannitol) were determined at pL 10.0 using protio or deuterio alcohol and were linear at the 95% confidence level. They revealed the independence of primary deuterium isotope effects on the atom fraction of deuterium in a mixed H(2)O-D(2)O solvent and yielded single-site transition-state fractionation factors of 0.43 +/- 0.05 and 0.47 +/- 0.01 for k(cat)/K(mannitol) and k(cat), respectively. (D)(k(cat)/K(mannitol)) was constant (1.80 +/- 0.20) in the pH range 6.0-9.5 and decreased at high pH to a limiting value of approximately 1. Measurement of (D)(k(cat)/K(fructose)) at pH 10.0 and 10.5 using NADH deuterium-labeled in the 4-pro-S position gave a value of 0.83, the equilibrium isotope effect on carbonyl group reduction. A mechanism of D-mannitol oxidation by MDH is supported by the data in which the partly rate-limiting transition state of hydride transfer is stabilized by a single solvation catalytic proton bridge. The chemical reaction involves a pH-dependent internal equilibrium which takes place prior to C-H bond cleavage and in which proton transfer from the reactive OH to the enzyme catalytic base may occur. Loss of a proton from the enzyme at high pH irreversibly locks the ternary complex with either alcohol or alkoxide bound in a conformation committed of undergoing NAD(+) reduction at a rate about 2.3-fold slower than the corresponding reaction rate of the protonated complex. Transient kinetic studies for D-mannitol oxidation at pH(D) 10.0 showed that the solvent isotope effect on steady-state turnover originates from a net rate constant of NADH release that is approximately 85% rate-limiting for k(cat) and 2-fold smaller in D(2)O than in H(2)O.  相似文献   

10.
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   

11.
We have investigated the steady state and equilibrium kinetic properties of carbonic anhydrase from Neisseria gonorrhoeae (NGCA). Qualitatively, the enzyme shows the same kinetic behaviour as the well studied human carbonic anhydrase II (HCA II). This is reflected in the similar pH dependencies of the kinetic parameters for CO(2) hydration and the similar behaviour of the kinetics of (18)O exchange between CO(2) and water at chemical equilibrium. The pH profile of the turnover number, k(cat), can be described as a titration curve with an exceptionally high maximal value of 1.7 x 10(6) s(-1) at alkaline pH and a pK(a) of 7.2. At pH 9, k(cat) is buffer dependent in a saturable manner, suggesting a ping-pong mechanism with buffer as the second substrate. The ratio k(cat)/K(m) is dependent on two ionizations with pK(a) values of 6.4 and 8.2. However, an (18)O-exchange assay identified only one ionizable group in the pH profile of k(cat)/K(m) with an apparent pK(a) of 6.5. The results of a kinetic analysis of a His66-->Ala variant of the bacterial enzyme suggest that His66 in NGCA has the same function as a proton shuttle as His64 in HCA II. The kinetic defect in the mutant can partially be overcome by certain buffers, such as imidazole and 1,2-dimethylimidazole. The bacterial enzyme shows similar K(i) values for the inhibitors NCO(-), SCN(-) and N(3)(-) as HCA II, while CN(-) and the sulfonamide ethoxzolamide are considerably weaker inhibitors of the bacterial enzyme than of HCA II. The absorption spectra of the adducts of Co(II)-substituted NGCA with acetazolamide, NCO(-), SCN(-), CN(-) and N(3)(-) resemble the corresponding spectra obtained with human Co(II)-isozymes I and II. Measurements of guanidine hydrochloride (GdnHCl)-induced denaturation reveal a sensitivity of the CO(2) hydration activity to the reducing agent tris(2-carboxyethyl)phosphine (TCEP). However, the A(292)/A(260) ratio was not affected by the presence of TCEP, and a structural transition at 2.8--2.9 M GdnHCl was observed.  相似文献   

12.
Mouse thymidylate synthase R209K (a mutation corresponding to R218K in Lactobacillus casei), overexpressed in thymidylate synthase-deficient Escherichia coli strain, was poorly soluble and with only feeble enzyme activity. The mutated protein, incubated with FdUMP and N(5,10)-methylenetetrahydrofolate, did not form a complex stable under conditions of SDS/polyacrylamide gel electrophoresis. The reaction catalyzed by the R209K enzyme (studied in a crude extract), compared to that catalyzed by purified wild-type recombinant mouse thymidylate synthase, showed the K(m) value for dUMP 571-fold higher and V(max) value over 50-fold (assuming that the mutated enzyme constituted 20% of total crude extract protein) lower. Thus the ratios k(cat, R209K)/k(cat, 'wild') and (k(cat, R209K)/K(m, R209K)(dUMP))/( k(cat, 'wild')/K(m, 'wild')(dUMP)) were 0.019 and 0.000032, respectively, documenting that mouse thymidylate synthase R209, similar to the corresponding L. casei R218, is essential for both dUMP binding and enzyme reaction.  相似文献   

13.
14.
Laccase from Myceliophthora thermophila (MtL) was expressed in functional form in Saccharomyces cerevisiae. Directed evolution improved expression eightfold to the highest yet reported for a laccase in yeast (18 mg/liter). Together with a 22-fold increase in kcat, the total activity was enhanced 170-fold. Specific activities of MtL mutants toward 2,2′-azino-bis(3-ethylbenzthiazoline-6-sulfonic acid) and syringaldazine indicate that substrate specificity was not changed by the introduced mutations. The most effective mutation (10-fold increase in total activity) introduced a Kex2 protease recognition site at the C-terminal processing site of the protein, adjusting the protein sequence to the different protease specificities of the heterologous host. The C terminus is shown to be important for laccase activity, since removing it by a truncation of the gene reduces activity sixfold. Mutations accumulated during nine generations of evolution for higher activity decreased enzyme stability. Screening for improved stability in one generation produced a mutant more stable than the heterologous wild type and retaining the improved activity. The molecular mass of MtL expressed in S. cerevisiae is 30% higher than that of the same enzyme expressed in M. thermophila (110 kDa versus 85 kDa). Hyperglycosylation, corresponding to a 120-monomer glycan on one N-glycosylation site, is responsible for this increase. This S. cerevisiae expression system makes MtL available for functional tailoring by directed evolution.  相似文献   

15.
Tsang A  Seidle H  Jawaid S  Zhou W  Smith C  Couch RD 《PloS one》2011,6(6):e20884
Deliberate and natural outbreaks of infectious disease, the prevalence of antibiotic resistant strains, and the ease by which antibiotic resistant bacteria can be intentionally engineered all underscore the necessity of effective vaccines and continued development of novel antimicrobial/antiviral therapeutics. Isoprenes, a group of molecules fundamentally involved in a variety of crucial biological functions, are derived from either the mevalonic acid (MVA) or methylerythritol phosphate (MEP) pathway. While mammals utilize the MVA pathway, many bacteria utilize the MEP pathway, highlighting the latter as an attractive target for antibiotic development. In this report we describe the cloning and characterization of Francisella tularensis MEP cytidylyltransferase, a MEP pathway enzyme and potential target for antibiotic development. Size exclusion chromatography indicates the protein exists as a dimer in solution. Enzyme assays produced an apparentK(MEP)(M) = 178 μM, K(CTP)(M) = 73 μM , k(MEP)(cat) = 1(s-1), k(CTP)(cat) = 0.8( s-1), and a k(MEP)(cat)/ K(MEP)(M) = 3.4 x 10(5) M(-1) min(-1). The enzyme exhibits a strict preference for Mg(+2) as a divalent cation and CTP as the nucleotide. Titanium dioxide chromatography-tandem mass spectrometry identified Thr141 as a site of phosphorylation. T141D and T141E site-directed mutants are catalytically inactive, suggesting a mechanism for post-translational control of metabolic flux through the F. tularensis MEP pathway. Overall, our study suggests that MEP cytidylyltransferase is an excellent target for the development of novel antibiotics against F. tularensis.  相似文献   

16.
The steady-state and pre-steady-state kinetic parameters for the interaction of E with the fluorogenic substrate 2-aminobenzoyl-Thr-Ile-Nle-Phe(p-NO(2))-Gln-Arg-NH(2) were determined in 1.25 M NaCl, 0.1 M MES-TRIS at pH 6.0 at 25 degrees C. At low concentrations of enzyme, the values of the K(m) and k(cat) calculated from steady-state data were 2.1 microM and 7.4 s(-1), respectively. At high concentrations of enzyme, the time-courses of fluorescence enhancement associated with catalysis were very dependent on the excitation wavelength used to monitor the reaction. Because the absorbance spectrum of the substrate overlapped the fluorescence emission spectrum of the enzyme, these abnormalities were attributed to fluorescence energy transfer between the enzyme and the substrate in an enzyme-substrate intermediate. The kinetic data collected with lambda(ex) = 280 nm and lambda(em) > 435 nm were analyzed according to the following mechanism in which EX was the species with enhanced fluorescence relative to substrate or products: [formula see text]. The values of the kinetic parameters with (1)H(2)O as the solvent were K = 13 microM, k(2) = 150 s(-1), k(-2) = 25 s(-1), and k(3) = 11 s(-1). The values of the kinetic parameters with (2)H(2)O as the solvent were K = 13 microM, k(2) = 210 s(-1), k(-2) = 12 s(-1), and k(3) = 4.4 s(-1). These values yielded solvent isotope effects of 2 on k(cat) and 0.9 on k(cat)/K(m). From analysis of the complete time-course of the fluorescence change (lambda(ex) = 280 nm and lambda(em) > 435 nm) during the course of substrate hydrolysis, the intermediate EX was determined to be 6.3-fold more fluorescent than the product, which, in turn, was 4.5-fold more fluorescent than ES or S. Rapid quench experiments with 2 N HCl as the quenching reagent confirmed that EX was a complex between enzyme and substrate. Consequently, the small burst in fluorescence observed when monitoring with lambda(ex) = 340 nm (0.3 product equiv per enzyme equivalent) was attributed to the fluorescence change upon transfer of substrate from an aqueous environment to a nonaqueous environment in the enzyme. These results were consistent with carbon-nitrogen bond cleavage being the major contributor to k(cat).  相似文献   

17.
Goto Y  Klinman JP 《Biochemistry》2002,41(46):13637-13643
Copper amine oxidases (CAOs) contain 2,4,5-trihydroxyphenylalanyl quinone (TPQ) and a copper ion in their active sites, catalyzing amine oxidation to aldehyde and ammonia concomitant with the reduction of molecular oxygen to hydrogen peroxide. Kinetic studies on the CAO from bovine serum (BSAO) [Su and Klinman (1999) Biochemistry 37, 12513-12525] and the recent reports on the cobalt substituted form of the enzyme from Hansenula polymorpha (HPAO) [Mills and Klinman (2000) J. Am. Chem. Soc. 122, 9897-9904, and Mills et al. (2002) Biochemistry, 41, 10577-10584] support pre-binding of molecular oxygen prior to a rate-limiting electron transfer from the reduced form of TPQ (p-aminohydroquinone form) to dioxygen. Although there is significant sequence homology between BSAO and HPAO, k(cat)/K(m)(O2) for BSAO under the optimal condition is one order of magnitude lower than that for HPAO. From a comparison of amino acid sequences for BSAO and HPAO, together with the X-ray crystal structure of HPAO, a plausible dioxygen pre-binding site has been identified that involves Y407, L425, and M634 in HPAO; the latter two residues are altered in BSAO to A490 and T695. To determine which of these residues plays a greater role in dioxygen chemistry, k(cat)/K(m)(O2) was determined in HPAO for the M634 --> T and L425 --> A mutants. The L425 --> A mutation does not alter k(cat)/K(m)(O2) to a large extent, whereas the M634 --> T decreased k(cat)/K(m)(O2) by one order of a magnitude, creating a catalyst that is similar to BSAO. A series of mutants at M634 (to F, L, and Q) were, therefore, prepared in HPAO and characterized with regard to k(cat)/K(m)(O2) as a function of pH. Structure reactivity correlations show a linear relationship of rate with side chain volume, rather than hydrophobicity, indicating that dioxygen reactivity increases with the bulk of the residue at position 634. This site also shows specificity for O2, in relation to the co-gas N2, since substitution of the inert gas N2 by either Ar or He has no effect on measured rates. In particular, He gas is expected to have little affinity for protein at 1 atmospheric pressure, implying little or no binding by N2 as well.  相似文献   

18.
Knapp MJ  Klinman JP 《Biochemistry》2003,42(39):11466-11475
The reactivity of O(2) with soybean lipoxygenase-1 (SLO) has been examined using a range of kinetic probes. We are able to rule out diffusional encounter of O(2) with protein, an outer-sphere electron transfer to O(2), and proton transfer as rate-limiting steps in k(cat)/K(M)(O(2)) for wild-type enzyme (WT SLO); this restricts the rate-limiting step to either the combination of O(2) with L(*) or a subsequent conformational change. In the Ile(553) --> Phe mutant, which constricts the putative O(2) binding channel [Knapp et al. (2001) J. Am. Chem. Soc. 123, 2931-2932], k(cat)/K(M)(O(2)) decreases by over a factor of 20; yet, this mutant appears to have the same rate-limiting step as WT SLO. It is argued that the slow step on k(cat)/K(M)(O(2)) is the combination of O(2) with L(*), with proximal protein effects determining the rate of reaction. The available data for SLO support the view that enzymes can affect O(2) reactivity without a direct involvement of metal cofactors. The primary role of the Fe(3+) cofactor is to generate an enzyme-bound radical, while the protein is concluded to control the stereo- and regiochemistry of O(2) encounter with this radical.  相似文献   

19.
To gain insight into the role of the strictly conserved histidine residue, H178, in the reaction mechanism of the methionyl aminopeptidase from Escherichia coli (EcMetAP-I), the H178A mutant enzyme was prepared. Metal-reconstituted H178A binds only one equivalent of Co(II) or Fe(II) tightly with affinities that are identical to the WT enzyme based on kinetic and isothermal titration calorimetry (ITC) data. Electronic absorption spectra of Co(II)-loaded H178A EcMetAP-I indicate that the active site divalent metal ion is pentacoordinate, identical to the WT enzyme. These data indicate that the metal binding site has not been affected by altering H178. The effect of altering H178 on activity is, in general, due to a decrease in k(cat). The k(cat) value for Co(II)-loaded H178A decreased 70-fold toward MGMM and 290-fold toward MP-p-NA compared to the WT enzyme, while k(cat) decreased 50-fold toward MGMM for the Fe(II)-loaded H178A enzyme and 140-fold toward MP-p-NA. The K(m) values for MGMM remained unaffected, while those for MP-p-NA increased approximately 2-fold for Co(II)- and Fe(II)-loaded H178A. The k(cat)/K(m) values for both Co(II)- and Fe(II)-loaded H178A toward both substrates ranged from approximately 50- to 580-fold reduction. The pH dependence of log K(m), log k(cat), and log(k(cat)/K(m)) of both WT and H178A EcMetAP-I were also obtained and are identical, within error, for H178A and WT EcMetAP-I. Therefore, H178A is catalytically important but is not required for catalysis. Assignment of one of the observed pK(a) values at 8.1 for WT EcMetAP-I was obtained from plots of molar absorptivity at lambda(max(640)) vs pH for both WT and H178A EcMetAP-I. Apparent pK(a) values of 8.1 and 7.6 were obtained for WT and H178A EcMetAP-I, respectively, and were assigned to the deprotonation of a metal-bound water molecule. The data reported herein provide support for the key elements of the previously proposed mechanism and suggest that a similar mechanism can apply to the enzyme with a single metal in the active site.  相似文献   

20.
The analysis of a multiple-sequence alignment (MSA) with correlation methods identifies pairs of residue positions whose occupation with amino acids changes in a concerted manner. It is plausible to assume that positions that are part of many such correlation pairs are important for protein function or stability. We have used the algorithm H2r to identify positions k in the MSAs of the enzymes anthranilate phosphoribosyl transferase (AnPRT) and indole-3-glycerol phosphate synthase (IGPS) that show a high conn(k) value, i.e., a large number of significant correlations in which k is involved. The importance of the identified residues was experimentally validated by performing mutagenesis studies with sAnPRT and sIGPS from the archaeon Sulfolobus solfataricus. For sAnPRT, five H2r mutant proteins were generated by replacing nonconserved residues with alanine or the prevalent residue of the MSA. As a control, five residues with conn(k) values of zero were chosen randomly and replaced with alanine. The catalytic activities and conformational stabilities of the H2r and control mutant proteins were analyzed by steady-state enzyme kinetics and thermal unfolding studies. Compared to wild-type sAnPRT, the catalytic efficiencies (k(cat)/K(M)) were largely unaltered. In contrast, the apparent thermal unfolding temperature (T(M)(app)) was lowered in most proteins. Remarkably, the strongest observed destabilization (ΔT(M)(app) = 14 °C) was caused by the V284A exchange, which pertains to the position with the highest correlation signal [conn(k) = 11]. For sIGPS, six H2r mutant and four control proteins with alanine exchanges were generated and characterized. The k(cat)/K(M) values of four H2r mutant proteins were reduced between 13- and 120-fold, and their T(M)(app) values were decreased by up to 5 °C. For the sIGPS control proteins, the observed activity and stability decreases were much less severe. Our findings demonstrate that positions with high conn(k) values have an increased probability of being important for enzyme function or stability.  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