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.
The genes (xylA) encoding xylose isomerase (XI) from two Lactococcus lactis subsp. lactis strains, 210 (Xyl) and IO-1 (Xyl+), were cloned, and the activities of their expressed proteins in recombinant strains of Escherichia coli were investigated. The nucleotide and amino acid sequence homologies between the xylA genes were 98.4 and 98.6%, respectively, and only six amino acid residues differed between the two XIs. The purified IO-1 XI was soluble with Km and kcat being 2.25 mM and 184/s, respectively, while the 210 XI was insoluble and inactive. Site-directed mutagenesis on 210 xylA showed that a triple mutant possessing R202M/Y218D/V275A mutations regained XI activity and was soluble. The Km and kcat of this mutant were 4.15 mM and 141/s, respectively. One of the IO-1 XI mutants, S388T, was insoluble and showed negligible activity similar to that of 210 XI. The introduction of a K407E mutation to the IO-1 S388T XI mutant restored its activity and solubility. The dissolution of XI activity in L. lactis subsp. lactis involves a series of mutations that collectively eliminate enzyme activity by reducing the solubility of the enzyme.  相似文献   

2.
Xylose metabolism, a variable phenotype in strains of Lactococcus lactis, was studied and evidence was obtained for the accumulation of mutations that inactivate the xyl operon. The xylose metabolism operon (xylRAB) was sequenced from three strains of lactococci. Fragments of 4.2, 4.2, and 5.4 kb that included the xyl locus were sequenced from L. lactis subsp. lactis B-4449 (formerly Lactobacillus xylosus), L. lactis subsp. lactis IO-1, and L. lactis subsp. lactis 210, respectively. The two environmental isolates, L. lactis B-4449 and L. lactis IO-1, produce active xylose isomerases and xylulokinases and can metabolize xylose. L. lactis 210, a dairy starter culture strain, has neither xylose isomerase nor xylulokinase activity and is Xyl(-). Xylose isomerase and xylulokinase activities are induced by xylose and repressed by glucose in the two Xyl(+) strains. Sequence comparisons revealed a number of point mutations in the xylA, xylB, and xylR genes in L. lactis 210, IO-1, and B-4449. None of these mutations, with the exception of a premature stop codon in xylB, are obviously lethal, since they lie outside of regions recognized as critical for activity. Nevertheless, either cumulatively or because of indirect affects on the structures of catalytic sites, these mutations render some strains of L. lactis unable to metabolize xylose.  相似文献   

3.
Tripeptidases from Lactococcus lactis subsp. lactis (L9PepTR), L. lactis subsp. cremoris (L6PepTR), and L. lactis subsp. hordniae (hTPepTR) were cloned, overexpressed, purified, and characterized. Although these enzymes contained three to seven naturally occurring amino acid differences, both metal-binding and catalytic sites were highly conserved. The k(cat) values of hTPepTR were approximately 1.5- to 2-fold higher than those of L9PepTR, while, for L6PepTR, they were approximately 0.8- to 1.4-times the L9PepTR values. The K(m) of tripeptidase from subsp. lactis (L9PepTR) was considerably larger when glycine was the amino acid located at both the N- and C-terminus of the peptide substrate. In addition, the K(m) values of L9PepTR increased in the following order for YGG, LGG, FGG, SGG, and alpha-aminoisobutyrylglycylglycine, while the k(cat)/K(m) decreased in the same order. These results suggest that the dipole moment and steric hindrance of the N-terminal amino acid side chain may be the most important factors controlling substrate specificity.  相似文献   

4.
Random PCR mutagenesis was applied to the Thermus thermophilus xylA gene encoding xylose isomerase. Three cold-adapted mutants were isolated with the following amino-acid substitutions: E372G, V379A (M-1021), E372G, F163L (M-1024) and E372G (M-1026). The wild-type and mutated xylA genes were cloned and expressed in Escherichia coli HB101 using the vector pGEM-T Easy, and their physicochemical and catalytic properties were determined. The optimum pH for xylose isomerization activity for the mutants was approximately 7.0, which is similar to the wild-type enzyme. Compared with the wild-type, the mutants were active over a broader pH range. The mutants exhibited up to nine times higher catalytic rate constants (k(cat)) for d-xylose compared with the wild-type enzyme at 60 degrees C, but they did not show any increase in catalytic efficiency (k(cat)/K(m)). For d-glucose, both the k(cat) and the k(cat)/K(m) values for the mutants were increased compared with the wild-type enzyme. Furthermore, the mutant enzymes exhibited up to 255 times higher inhibition constants (K(i)) for xylitol than the wild-type, indicating that they are less inhibited by xylitol. The thermal stability of the mutated enzymes was poorer than that of the wild-type enzyme. The results are discussed in terms of increased molecular flexibility of the mutant enzymes at low temperatures.  相似文献   

5.
We report the complete genome sequence of Lactococcus lactis IO-1 (= JCM7638). It is a nondairy lactic acid bacterium, produces nisin Z, ferments xylose, and produces predominantly L-lactic acid at high xylose concentrations. From ortholog analysis with other five L. lactis strains, IO-1 was identified as L. lactis subsp. lactis.  相似文献   

6.
Four mutations observed between tripeptidases from Lactococcus lactis subsp. lactis and subsp. cremoris were introduced one by one to the corresponding points in wild-type tripeptidase from L. lactis subsp. lactis. The k(cat) values of four resultant mutants were analyzed and discussed in stereographical terms. Change in catalytic activity appeared to be related to the sequential and steric location of mutation point within the enzyme protein, even though no drastic change was observed with one point mutation.  相似文献   

7.
A direct mass spectrometric approach was used for the determination of steady-state kinetic parameters, the turnover number (k(cat)), the Michaelis constant (K(M)), and the specificity constant (k(cat)/K(M)) for an enzyme-catalyzed hydrolysis of xylooligosaccharides. Electrospray ionization mass spectrometry was performed to observe product distributions and to determine k(cat), K(M), and k(cat)/K(M) values for Trichoderma reesei endo-1,4-beta-xylanase II (TRX II) with xylohexaose (Xyl(6)), xylopentaose (Xyl(5)), xylotetraose (Xyl(4)), and xylotriose (Xyl(3)) as substrates. The determined k(cat)/K(M) values (0.93, 0.37, 0.027, and 0.00015 microM(-1) s(-1), respectively) indicated that Xyl(6) was the most preferred substrate of TRX II. In addition, the obtained K(M) value for Xyl(5) (136 microM) was roughly twice as high as that for Xyl(6) (73 microM), suggesting that at least six putative subsites contribute to the substrate binding in the active site of TRX II. Previous mass spectrometric assays for enzyme kinetics have been used mostly in the case of reactions that result in a transfer of acidic groups (e.g., phosphate) into neutral oligosaccharides giving rise to negatively charged products. Here we demonstrate that such analysis is also feasible in the case of neutral underivatized oligosaccharides. Implications of the results for the catalytic mechanism of TRX II in particular are discussed.  相似文献   

8.
Genetic and biochemical evidence for a defective xylan degradation pathway was found linked to the xylose operon in three lactococcal strains, Lactococcus lactis 210, L. lactis IO-1, and L. lactis NRRL B-4449. Immediately downstream of the xylulose kinase gene (xylB) (K. A. Erlandson, J.-H. Park, W. El Khal, H.-H. Kao, P. Basaran, S. Brydges, and C. A. Batt, Appl. Environ. Microbiol. 66:3974-3980, 1999) are two open reading frames encoding a mutarotase (xylM) and a xyloside transporter (xynT) and a partial open reading frame encoding a beta-xylosidase (xynB). These are functions previously unreported for lactococci or lactobacilli. The mutarotase activity of the putative xylM gene product was confirmed by overexpression of the L. lactis enzyme in Escherichia coli and purification of recombinant XylM. We hypothesize that the mutarotase links xylan degradation to xylose metabolism due to the anomeric preference of xylose isomerase. In addition, Northern hybridization experiments suggested that the xylM and xynTB genes are cotranscribed with the xylRAB genes, responsible for xylose metabolism. Although none of the three strains appeared to metabolize xylan or xylobiose, they exhibited xylosidase activity, and L. lactis IO-1 and L. lactis NRRL B-4449 had functional mutarotases.  相似文献   

9.
A spontaneous derivative of Lactococcus lactis subsp. lactis K1 (formerly Streptococcus lactis K1) lacking N5-(carboxyethyl)ornithine synthase (EC 1.5.1.24) was isolated. This mutant had also lost the abilities to ferment sucrose and to produce the antibiotic nisin. Hybridization studies indicate that these linked traits are encoded on the chromosome of L. lactis K1 and that they may be located on a conjugative transposon.  相似文献   

10.
We describe here a new enzyme-coupled assay for the quantitation of d-xylose using readily available enzymes that allows kinetic evaluation of hemicellulolytic enzymes using natural xylooligosaccharide substrates. Hydrogen peroxide is generated as an intermediary analyte, which allows flexibility in the choice of the chromophore or fluorophore used as the final reporter. Thus, we present d-xylose quantitation results for solution-phase assays performed with both the fluorescent reporter resorufin, generated from N-acetyl-3,7-dihydroxyphenoxazine (Amplex Red), and 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulfonate) (ABTS), whose corresponding radical cation has an absorbance maximum at approximately 400 nm. We also describe a useful solid-phase variation of the assay performed with the peroxidase substrate 3,3'-diaminobenzidine tetrahydrochloride, which produces an insoluble brown precipitate. In addition, kinetic parameters for hydrolysis of the natural substrates xylobiose and xylotriose were obtained using this assay for a glycosyl hydrolase family 39 beta-xylosidase from Thermoanaerobacterium sp. strain JW/SL YS485 (Swiss-Prot accession no. O30360). At higher xylobiose substrate concentrations the enzyme showed an increase in the rate indicative of transglycosylation, while for xylotriose marked substrate inhibition was observed. At lower xylobiose concentrations k(cat) was 2.7 +/- 0.4 s(-1), K(m) was 3.3 +/- 0.7 mM, and k(cat)/K(m) was 0.82 +/- 0.21 mM(-1) . s(-1). Nonlinear curve fitting to a substrate inhibition model showed that for xylotriose K(i) was 1.7 +/- 0.1 mM, k(cat) was 2.0 +/- 0.1 s(-1), K(m) was 0.144 +/- 0.011 mM, and k(cat)/K(m) was 14 +/- 1.3 mM(-1) . s(-1).  相似文献   

11.
An uncharacterized gene from Thermus thermophilus, thought to encode a mannose-6-phosphate isomerase, was cloned and expressed in Escherichia coli. The maximal activity of the recombinant enzyme for L-ribulose isomerization was observed at pH 7.0 and 75°C in the presence of 0.5 mM Cu(2+). Among all of the pentoses and hexoses evaluated, the enzyme exhibited the highest activity for the conversion of L-ribulose to L-ribose, a potential starting material for many L-nucleoside-based pharmaceutical compounds. The active-site residues, predicted according to a homology-based model, were separately replaced with Ala. The residue at position 142 was correlated with an increase in L-ribulose isomerization activity. The R142N mutant showed the highest activity among mutants modified with Ala, Glu, Tyr, Lys, Asn, or Gln. The specific activity and catalytic efficiency (k(cat)/K(m)) for L-ribulose using the R142N mutant were 1.4- and 1.6-fold higher than those of the wild-type enzyme, respectively. The k(cat)/K(m) of the R142N mutant was 3.8-fold higher than that of Geobacillus thermodenitrificans mannose-6-phosphate isomerase, which exhibited the highest activity to date for the previously reported k(cat)/K(m). The R142N mutant enzyme produced 213 g/liter L-ribose from 300 g/liter L-ribulose for 2 h, with a volumetric productivity of 107 g liter(-1) h(-1), which was 1.5-fold higher than that of the wild-type enzyme.  相似文献   

12.
β-Xylosidases are hemicellulases that hydrolyze short xylo-oligosaccharides into xylose units, thus complementing endoxylanase degradation of the hemicellulose component of lignocellulosic substrates. Here, we describe the cloning, characterization, and kinetic analysis of a glycoside hydrolase family 43 β-xylosidase (Xyl43A) from the aerobic cellulolytic bacterium, Thermobifida fusca. Temperature and pH optima of 55-60 °C and 5.5-6, respectively, were determined. The apparent K(m) value was 0.55 mM, using p-nitrophenyl xylopyranoside as substrate, and the catalytic constant (k(cat)) was 6.72 s(-1). T. fusca Xyl43A contains a catalytic module at the N terminus and an ancillary module (termed herein as Module-A) of undefined function at the C terminus. We expressed the two recombinant modules independently in Escherichia coli and examined their remaining catalytic activity and binding properties. The separation of the two Xyl43A modules caused the complete loss of enzymatic activity, whereas potent binding to xylan was fully maintained in the catalytic module and partially in the ancillary Module-A. Nondenaturing gel electrophoresis revealed a specific noncovalent coupling of the two modules, thereby restoring enzymatic activity to 66.7% (relative to the wild-type enzyme). Module-A contributes a phenylalanine residue that functions as an essential part of the active site, and the two juxtaposed modules function as a single functional entity.  相似文献   

13.
Mu W  Yu S  Jiang B  Li X 《Biotechnology letters》2012,34(5):907-911
The gene coding for D-lactate dehydrogenase (D-LDH) from Pediococcus acidilactici DSM 20284 was cloned and expressed in E. coli. The recombinant enzyme was purified by nickel-affinity chromatography. It converted phenylpyruvic acid (PPA) to 3-phenyllactic acid maximally at 30°C and pH 5.5 with a specific activity of 140 and 422 U/mg for PPA and pyruvate, respectively. The K(m), turnover number (k(cat)), and catalytic efficiency (k(cat)/K(m)) for PPA were 2.9 mM, 305 s(-1), and 105 mM(-1) s(-1), respectively.  相似文献   

14.
Extracellular Penicillium fellutanum exo-beta-D-galactofuranosidase, with a mass of 70 kDa, was purified to apparent homogeneity. The enzyme was used to investigate the influence of phosphodiesters of the peptidophosphogalactomannans pP(2)GM(ii) and pP(25)GM(ii) (containing 2 and 25 phosphodiester residues, respectively, per mol of polymer) on the kinetic parameters of galactofuranosyl hydrolysis of these two polymers, of 1-O-methyl-beta-D-galactofuranoside, and of two galactofuranooligosaccharides. The enzyme did not hydrolyze phosphorylated galactose residues of pP(2)GM(ii) or pP(25)GM(ii). The k(cat)/K(m) value for pP(25)GM(ii) is 1.7 x 10(3) M(-1) s(-1), that for 1-O-methyl-beta-D-galactofuranoside is 1.1 x 10(4) M(-1) s(-1), that for pP(2)GM(ii) is 1.7 x 10 (4) M(-1) s(-1), and those for 5-O-beta-D-galactofuranooligosaccharides with degrees of polymerization of 3.4 and 5.5 are 1.7 x 10(5) and 4.1 x 10(5) M(-1) s(-1), respectively. Variability in the k(cat)/K(m) values is due primarily to differences in K(m) values; the k(-1)/k(1) ratio likely provides the most influence on K(m). k(cat) increases as the degree of polymerization of galactofuranosyl residues increases. Most of the galactofuranosyl and phosphocholine residues were removed by day 8 in vivo from pP(x)GM(ii) added to day 3 cultures initiated in medium containing 2 mM phosphate but not from those initially containing 20 mM phosphate. The filtrates from day 9 cultures initiated in 2 mM inorganic phosphate in modified Raulin-Thom medium contained 0.2 mM inorganic phosphate and 2.2 U of galactofuranosidase ml(-1)h(-1). No galactofuranosidase activity but 15 mM inorganic phosphate was found in filtrates from day 9 cultures initiated in 20 mM phosphate. In vivo the rate of galactofuranosyl hydrolysis of pP(x)GM(ii) and of related polymers is proportional to the k(cat)/K(m) value of each polymer. The kinetic data show that the k(cat)/K(m) value increases as the number of phosphodiesters of pP(x)GM(ii) decreases, also resulting in an increase in the activity of exo-beta-D-galactofuranosidase.  相似文献   

15.
A highly thermostable xylanase (Xyl I) produced by Thermomonospora sp. was purified to homogeneity and was classified as a family 10 xylanase based on its molecular weight (38,000 Da) and isoelectric point (4.1). K2d analysis showed that the secondary structure of Xyl I was made up of 38% alpha-helix and 10% beta-sheet. The optimal temperature for the activity of Xyl I was 80 degrees C. Xyl I was highly thermostable with half-lives of 86, 30, and 15 min at 80, 90, and 100 degrees C respectively. Xyl I was stable in an expansive pH range of 5 to 10 with more than 75% residual activity. Our present investigation using o-phthalaldehyde (OPTA) as the chemical initiator for fluorescent chemoaffinity labeling and trinitrobenzenesulphonic acid (TNBS) as chemical modifier have revealed the presence of a single lysine residue in the active site of Xyl I. The high pK value for the basic limb of the pH profile reflects the ionization of a lysine residue. The higher K(m) values and similar k(cat) values of the TNBS modified enzyme in comparison to native enzyme and the substrate protection against OPTA and TNBS, suggested the presence of the lysine residue in the substrate-binding site.  相似文献   

16.
A chimeric gene, Glu-Xyl, encoding Bacillus amyloliquefaciens glucanase (Glu, 24.4 kDa) and Bacillus subtilis xylanase (Xyl, 21.2 kDa), was constructed via end-to-end fusion and expressed successfully in Escherichia coli. The purified fusion protein (46.1 kDa) exhibited both glucanase and xylanase activities. Compared with parental enzymes, the Glu moiety was characterized by kinetic parameters of decreased K(m) (0.66-fold) and increased K(cat) (2.75-fold), whereas the Xyl moiety had an increased K(m) (1.37-fold) and decreased K(cat) (0.79-fold). These indicate a 3.15-fold net increase and a 31% decrease in catalytic efficiency (K(cat)/K(m)) of the Glu and Xyl moieties. Activities and stabilities of both moieties at 40-90 degrees C or pH 3.0-10.0 were compared with those of the parental enzymes. Despite some variations, common optima were 40 degrees C and pH 9.0 for the Glu moiety and parent, and 50-60 degrees C and pH 9.0 for the Xyl counterparts. Thus, the fusion enzyme Glu-Xyl was bifunctional, with greatly enhanced glucanase activity associated with a decrease in xylanase activity.  相似文献   

17.
A simple and direct assay method for glucose oxidase (EC 1.1.3.4) from Aspergillus niger and Penicillium amagasakiense was investigated by Fourier transform infrared spectroscopy. This enzyme catalyzed the oxidation of d-glucose at carbon 1 into d-glucono-1,5-lactone and hydrogen peroxide in phosphate buffer in deuterium oxide ((2)H(2)O). The intensity of the d-glucono-1,5-lactone band maximum at 1212 cm(-1) due to CO stretching vibration was measured as a function of time to study the kinetics of d-glucose oxidation. The extinction coefficient epsilon of d-glucono-1,5-lactone was determined to be 1.28 mM(-1)cm(-1). The initial velocity is proportional to the enzyme concentration by using glucose oxidase from both A. niger and P. amagasakiense either as cell-free extracts or as purified enzyme preparations. The kinetic constants (V(max), K(m), k(cat), and k(cat)/K(m)) determined by Lineweaver-Burk plot were 433.78+/-59.87U mg(-1) protein, 10.07+/-1.75 mM, 1095.07+/-151.19s(-1), and 108.74 s(-1)mM(-1), respectively. These data are in agreement with the results obtained by a spectrophotometric method using a linked assay based on horseradish peroxidase in aqueous media: 470.36+/-42.83U mg(-1) protein, 6.47+/-0.85 mM, 1187.77+/-108.16s(-1), and 183.58 s(-1)mM(-1) for V(max), K(m), k(cat), and k(cat)/K(m), respectively. Therefore, this spectroscopic method is highly suited to assay for glucose oxidase activity and its kinetic parameters by using either cell-free extracts or purified enzyme preparations with an additional advantage of performing a real-time measurement of glucose oxidase activity.  相似文献   

18.
Wood MJ  Becvar LA  Prieto JH  Melacini G  Komives EA 《Biochemistry》2003,42(41):11932-11942
Oxidation of Met 388, one of the three linker residues connecting the fourth and fifth EGF-like domains of thrombomodulin (TM), is deleterious for TM activity. An NMR structure of the smallest active fragment of TM (TMEGF45) and a crystal structure of a larger fragment (TMEGF456) bound to thrombin both show that Met 388 is packed into the fifth domain. Using multidimensional NMR, we have solved the structure of TMEGF45 in which Met 388 is oxidized (TMEGF45ox) and the structure of TMEGF45 in which Met 388 is mutated to Leu (TMEGF45ML). Comparison of the structures shows that the fifth domain has a somewhat different structure depending on the residue at position 388, and several of the thrombin-binding residues are packed into the fifth domain in the oxidized protein while they are exposed and free to interact with thrombin in the native structure and the Met-Leu mutant. This observation is consistent with kinetic measurements showing that the K(m) for TMEGF45ox binding to thrombin is 3.3-fold higher than for the native protein. Most importantly, the connection between the two domains, as indicated by interdomain NOEs, appears to be essential for activity. In the TMEGF45ox structure which has a reduced k(cat) for protein C activation by the thrombin-TMEGF45ox complex, interaction between the two domains is lost. Conversely, a tighter connection is observed between the two domains in TMEGF45ML, which has a higher k(cat) for protein C activation by the thrombin-TMEGF45ML complex.  相似文献   

19.
Toteva MM  Silvaggi NR  Allen KN  Richard JP 《Biochemistry》2011,50(46):10170-10181
D-Xylose isomerase (XI) and triosephosphate isomerase (TIM) catalyze the aldose-ketose isomerization reactions of D-xylose and d-glyceraldehyde 3-phosphate (DGAP), respectively. D-Glyceraldehyde (DGA) is the triose fragment common to the substrates for XI and TIM. The XI-catalyzed isomerization of DGA to give dihydroxyacetone (DHA) in D(2)O was monitored by (1)H nuclear magnetic resonance spectroscopy, and a k(cat)/K(m) of 0.034 M(-1) s(-1) was determined for this isomerization at pD 7.0. This is similar to the k(cat)/K(m) of 0.017 M(-1) s(-1) for the TIM-catalyzed carbon deprotonation reaction of DGA in D(2)O at pD 7.0 [Amyes, T. L., O'Donoghue, A. C., and Richard, J. P. (2001) J. Am. Chem. Soc. 123, 11325-11326]. The much larger activation barrier for XI-catalyzed isomerization of D-xylose (k(cat)/K(m) = 490 M(-1) s(-1)) versus that for the TIM-catalyzed isomerization of DGAP (k(cat)/K(m) = 9.6 × 10(6) M(-1) s(-1)) is due to (i) the barrier to conversion of cyclic d-xylose to the reactive linear sugar (5.4 kcal/mol) being larger than that for conversion of DGAP hydrate to the free aldehyde (1.7 kcal/mol) and (ii) the intrinsic binding energy [Jencks, W. P. (1975) Adv. Enzymol. Relat. Areas Mol. Biol. 43, 219-410] of the terminal ethylene glycol fragment of D-xylose (9.3 kcal/mol) being smaller than that of the phosphodianion group of DGAP (~12 kcal/mol). The XI-catalyzed isomerization of DGA in D(2)O at pD 7.0 gives a 90% yield of [1-(1)H]DHA and a 10% yield of [1-(2)H]DHA, the product of isomerization with incorporation of deuterium from solvent D(2)O. By comparison, the transfer of (3)H from the labeled hexose substrate to solvent is observed only once in every 10(9) turnovers for the XI-catalyzed isomerization of [2-(3)H]glucose in H(2)O [Allen, K. N., Lavie, A., Farber, G. K., Glasfeld, A., Petsko, G. A., and Ringe, D. (1994) Biochemistry 33, 1481-1487]. We propose that truncation of the terminal ethylene glycol fragment of d-xylose to give DGA results in a large decrease in the rate of XI-catalyzed isomerization with hydride transfer compared with that for proton transfer. An ultra-high-resolution (0.97 ?) X-ray crystal structure was determined for the complex obtained by soaking crystals of XI with 50 mM DGA. The triose binds to XI as the unreactive hydrate, but ligand binding induces metal cofactor movement and conformational changes in active site residues similar to those observed for XI·sugar complexes.  相似文献   

20.
The roles of three conserved active site carboxylic acids (D197, E233, and D300) in the catalytic mechanism of human pancreatic alpha-amylase (HPA) were studied by utilizing site-directed mutagenesis in combination with structural and kinetic analyses of the resultant enzymes. All three residues were mutated to both alanine and the respective amide, and a double alanine mutant (E233A/D300A) was also generated. Structural analyses demonstrated that there were no significant differences in global fold for the mutant enzymes. Kinetic analyses were performed on the mutants, utilizing a range of substrates. All results suggested that D197 was the nucleophile, as virtually all activity (>10(5)-fold decrease in k(cat) values) was lost for the enzymes mutated at this position when assayed with several substrates. The significantly greater second-order rate constant of E233 mutants on "activated" substrates (k(cat)/K(m) value for alpha-maltotriosyl fluoride = 15 s(-)(1) mM(-)(1)) compared with "unactivated" substrates (k(cat)/K(m) value for maltopentaose = 0.0030 s(-)(1) mM(-)(1)) strongly suggested that E233 is the general acid catalyst, as did the pH-activity profiles. Transglycosylation was favored over hydrolysis for the reactions of several of the enzymes mutated at D300. At the least, this suggests an overall impairment of the catalytic mechanism where the reaction then proceeds using the better acceptor (oligosaccharide instead of water). This may also suggest that D300 plays a crucial role in enzymic interactions with the nucleophilic water during the hydrolysis of the glycosidic bond.  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