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.
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   

2.
Cellobiose dehydrogenase is an extracellular flavocytochrome, which catalyzes the oxidation of cellobiose and other soluble oligosaccharides to their respective lactones, while reducing various one- and two-electron acceptors. Two residues at the active site of the flavin domain, His689 and Asn732, have been proposed to play critical roles in the oxidation of the substrate. To test these proposals, each residue was substituted with either a Gln, Asn, Glu, Asp, Val, Ala, and/or a His residue by site-directed mutagenesis, using a homologous expression system previously developed in our laboratory. This enabled an examination of the functional, stereochemical, and electrostatic constraints for binding and oxidation of the substrate. The steady-state kinetic parameters for the variant proteins were compared using cellobiose and its epimer, lactose, as the substrates. The H689 variants all exhibit >1000-fold lower k(cat) values, while the K(m) values for both substrates in these variants are similar to that of the wild-type enzyme. This supports the proposed role of this His residue as a general base in catalysis. The N732 variants exhibit a range of kinetic parameters: the k(cat) values for oxidation are 5-4000-fold lower than that for the wild-type enzyme, while the K(m) values vary between similar to and 60-fold higher than that for the wild-type. The difference in binding energy between cellobiose and lactose was calculated using the relationship delta(delta G) = -RT ln[(k(cat)/K(m))(lactose)/(k(cat)/K(m))(cellobiose)]. This calculation for the wild-type enzyme suggests that lactose binds considerably more weakly than cellobiose (7.2 kJ/mol difference), which corresponds to one extra (cumulative) hydrogen bond for cellobiose over lactose. Mutations at Asn732 result in a further weakening of lactose binding over cellobiose (2-4 kJ/mol difference). The results support a role for Asn732 in the binding of the substrate.  相似文献   

3.
Glutathione transferase (GST) A3-3 is the most efficient human steroid double-bond isomerase known. The activity with Delta(5)-androstene-3,17-dione is highly dependent on the phenolic hydroxyl group of Tyr-9 and the thiolate of glutathione. Removal of these groups caused an 1.1 x 10(5)-fold decrease in k(cat); the Y9F mutant displayed a 150-fold lower isomerase activity in the presence of glutathione and a further 740-fold lower activity in the absence of glutathione. The Y9F mutation in GST A3-3 did not markedly decrease the activity with the alternative substrate 1-chloro-2,4-dinitrobenzene. Residues Phe-10, Leu-111, and Ala-216 selectively govern the activity with the steroid substrate. Mutating residue 111 into phenylalanine caused a 25-fold decrease in k(cat)/K(m) for the steroid isomerization. The mutations A216S and F10S, separate or combined, affected the isomerase activity only marginally, but with the additional L111F mutation k(cat)/K(m) was reduced to 0.8% of that of the wild-type value. In contrast, the activities with 1-chloro-2,4-dinitrobenzene and phenethylisothiocyanate were not largely affected by the combined mutations F10S/L111F/A216S. K(i) values for Delta(5)-androstene-3,17-dione and Delta(4)-androstene-3,17-dione were increased by the triple mutation F10S/L111F/A216S. The pK(a) of the thiol group of active-site-bound glutathione, 6.1, increased to 6.5 in GST A3-3/Y9F. The pK(a) of the active-site Tyr-9 was 7.9 for the wild-type enzyme. The pH dependence of k(cat)/K(m) of wild-type GST A3-3 for the isomerase reaction displays two kinetic pK(a) values, 6.2 and 8.1. The basic limb of the pH dependence of k(cat) and k(cat)/K(m) disappears in the Y9F mutant. Therefore, the higher kinetic pK(a) reflects ionization of Tyr-9, and the lower one reflects ionization of glutathione. We propose a reaction mechanism for the double-bond isomerization involving abstraction of a proton from C4 in the steroid accompanied by protonation of C6, the thiolate of glutathione serving as a base and Tyr-9 assisting by polarizing the 3-oxo group of the substrate.  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.
L-Ribulose is a rare and expensive sugar that can be used as a precursor for the production of other rare sugars of high market value such as L-ribose. In this work we describe a production process for L-ribulose using L-arabinose, a common component of polymers of lignocellulosic materials, as the starting material. A ribulokinase-deficient mutant of the heterofermentative lactic acid bacterium Lactobacillus plantarum NCIMB8826 was constructed. Expression of araA, which encodes the critical enzyme L-arabinose isomerase, was repressed by high glucose concentrations in batch cultivations. A fed-batch cultivation strategy was therefore used to maximize L-arabinose isomerase production during growth. Resting cells of the ribulokinase-deficient mutant were used for the production of L-ribulose. The isomerization of L-arabinose to L-ribulose was very unfavorable for L-ribulose formation. However, high L-ribulose yields were obtained by complexing the produced L-ribulose with borate. The process for L-ribulose production in borate buffer by resting cells was optimized using central composite designs. The experiment design suggested that the process has an optimal operation point around an L-arabinose concentration of 100 g liter(-1), a borate concentration of 500 mM, and a temperature of 48 degrees C, where the statistical software predicted an initial L-ribulose production rate of 29.1 g liter(-1) h(-1), a best-achievable process productivity of 14.8 g liter(-1) h(-1), and a conversion of L-arabinose to L-ribulose of 0.70 mol mol(-1).  相似文献   

6.
Pichia stipitis NAD(+)-dependent xylitol dehydrogenase (XDH), a medium-chain dehydrogenase/reductase, is one of the key enzymes in ethanol fermentation from xylose. For the construction of an efficient biomass-ethanol conversion system, we focused on the two areas of XDH, 1) change of coenzyme specificity from NAD(+) to NADP(+) and 2) thermostabilization by introducing an additional zinc atom. Site-directed mutagenesis was used to examine the roles of Asp(207), Ile(208), Phe(209), and Asn(211) in the discrimination between NAD(+) and NADP(+). Single mutants (D207A, I208R, F209S, and N211R) improved 5 approximately 48-fold in catalytic efficiency (k(cat)/K(m)) with NADP(+) compared with the wild type but retained substantial activity with NAD(+). The double mutants (D207A/I208R and D207A/F209S) improved by 3 orders of magnitude in k(cat)/K(m) with NADP(+), but they still preferred NAD(+) to NADP(+). The triple mutant (D207A/I208R/F209S) and quadruple mutant (D207A/I208R/F209S/N211R) showed more than 4500-fold higher values in k(cat)/K(m) with NADP(+) than the wild-type enzyme, reaching values comparable with k(cat)/K(m) with NAD(+) of the wild-type enzyme. Because most NADP(+)-dependent XDH mutants constructed in this study decreased the thermostability compared with the wild-type enzyme, we attempted to improve the thermostability of XDH mutants by the introduction of an additional zinc atom. The introduction of three cysteine residues in wild-type XDH gave an additional zinc-binding site and improved the thermostability. The introduction of this mutation in D207A/I208R/F209S and D207A/I208R/F209S/N211R mutants increased the thermostability and further increased the catalytic activity with NADP(+).  相似文献   

7.
Tripp BC  Ferry JG 《Biochemistry》2000,39(31):9232-9240
Four glutamate residues in the prototypic gamma-class carbonic anhydrase from Methanosarcina thermophila (Cam) were characterized by site-directed mutagenesis and chemical rescue studies. Alanine substitution indicated that an external loop residue, Glu 84, and an internal active site residue, Glu 62, are both important for CO(2) hydration activity. Two other external loop residues, Glu 88 and Glu 89, are less important for enzyme function. The two E84D and -H variants exhibited significant activity relative to wild-type activity in pH 7.5 MOPS buffer, suggesting that the original glutamate residue could be substituted with other ionizable residues with similar pK(a) values. The E84A, -C, -K, -Q, -S, and -Y variants exhibited large decreases in k(cat) values in pH 7.5 MOPS buffer, but only exhibited small changes in k(cat)/K(m). These same six variants were all chemically rescued by pH 7.5 imidazole buffer, with 23-46-fold increases in the apparent k(cat). These results are consistent with Glu 84 functioning as a proton shuttle residue. The E62D variant exhibited a 3-fold decrease in k(cat) and a 2-fold decrease in k(cat)/K(m) relative to those of the wild type in pH 7.5 MOPS buffer, while other substitutions (E62A, -C, -H, -Q, -T, and -Y) resulted in much larger decreases in both k(cat) and k(cat)/K(m). Imidazole did not significantly increase the k(cat) values and slightly decreased the k(cat)/K(m) values of most of the Glu 62 variants. These results indicate a primary preference for a carboxylate group at position 62, and support a proposed catalytic role for residue Glu 62 in the CO(2) hydration step, but do not definitively establish its role in the proton transport step.  相似文献   

8.
The functional role of the highly conserved active site Arg 59 in the prototype of the gamma-class carbonic anhydrase Cam (carbonic anhydrase from Methanosarcina thermophila) was investigated. Variants (R59A, -C, -E, -H, -K, -M, and -Q) were prepared by site-directed mutagenesis and characterized by size exclusion chromatography (SEC), circular dichroism (CD) spectroscopy, and stopped-flow kinetic analyses. CD spectra indicated similar secondary structures for the wild type and the R59A and -K variants, independent of nondenaturing concentrations of guanidine hydrochloride (GdnHCl). SEC indicated that all variants purified as homotrimers like the wild type. SEC also revealed that the R59A and -K variants unfolded at > or = 1.5 M GdnHCl, compared to 3.0 M GdnHCl for the wild type. These results indicate that Arg 59 contributes to the thermodynamic stability of the Cam trimer. The R59K variant had k(cat) and k(cat)/K(m) values that were 8 and 5% of the wild-type values, respectively, while all other variants had k(cat) and k(cat)/K(m) values 10-100-fold lower than those of the wild type. The R59A, -C, -E, -M, and -Q variants exhibited 4-63-fold increases in k(cat) and 9-120-fold increases in k(cat)/K(m) upon addition of 100 mM GdnHCl, with the largest increases observed for the R59A variant, which was comparable to the R59K variant. The kinetic results indicate that a positive charge at position 59 is essential for the CO(2) hydration step of the overall catalytic mechanism.  相似文献   

9.
Klebsiella aerogenes urease uses a dinuclear nickel active site to catalyze urea hydrolysis at >10(14)-fold the spontaneous rate. To better define the enzyme mechanism, we examined the kinetics and structures for a suite of site-directed variants involving four residues at the active site: His320, His219, Asp221, and Arg336. Compared to wild-type urease, the H320A, H320N, and H320Q variants exhibit similar approximately 10(-)(5)-fold deficiencies in rates, modest K(m) changes, and disorders in the peptide flap covering their active sites. The pH profiles for these mutant enzymes are anomalous with optima near 6 and shoulders that extend to pH 9. H219A urease exhibits 10(3)-fold increased K(m) over that of native enzyme, whereas the increase is less marked ( approximately 10(2)-fold) in the H219N and H219Q variants that retain hydrogen bonding capability. Structures for these variants show clearly resolved active site water molecules covered by well-ordered peptide flaps. Whereas the D221N variant is only moderately affected compared to wild-type enzyme, D221A urease possesses low activity ( approximately 10(-)(3) that of native enzyme), a small increase in K(m), and a pH 5 optimum. The crystal structure for D221A urease is reminiscent of the His320 variants. The R336Q enzyme has a approximately 10(-)(4)-fold decreased catalytic rate with near-normal pH dependence and an unaffected K(m). Phenylglyoxal inactivates the R336Q variant at over half the rate observed for native enzyme, demonstrating that modification of non-active-site arginines can eliminate activity, perhaps by affecting the peptide flap. Our data favor a mechanism in which His219 helps to polarize the substrate carbonyl group, a metal-bound terminal hydroxide or bridging oxo-dianion attacks urea to form a tetrahedral intermediate, and protonation occurs via the general acid His320 with Asp221 and Arg336 orienting and influencing the acidity of this residue. Furthermore, we conclude that the simple bell-shaped pH dependence of k(cat) and k(cat)/K(m) for the native enzyme masks a more complex underlying pH dependence involving at least four pK(a)s.  相似文献   

10.
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   

11.
The roles of an aspartate and an arginine, which are completely conserved in the active sites of beta-class carbonic anhydrases, were investigated by steady-state kinetic analyses of replacement variants of the beta-class enzyme (Cab) from the archaeon Methanobacterium thermoautotrophicum. Previous kinetic analyses of wild-type Cab indicated a two-step zinc-hydroxide mechanism of catalysis in which the k(cat)/K(m) value depends only on the rate constants for the CO(2) hydration step, whereas k(cat) also depends on rate constants from the proton transfer step (K. S. Smith, N. J. Cosper, C. Stalhandske, R. A. Scott, and J. G. Ferry, J. Bacteriol. 182:6605-6613, 2000). The recently solved crystal structure of Cab shows the presence of a buffer molecule within hydrogen bonding distance of Asp-34, implying a role for this residue in the proton transport step (P. Strop, K. S. Smith, T. M. Iverson, J. G. Ferry, and D. C. Rees, J. Biol. Chem. 276:10299-10305, 2001). The k(cat)/K(m) values of Asp-34 variants were decreased relative to those of the wild type, although not to an extent which supports an essential role for this residue in the CO(2) hydration step. Parallel decreases in k(cat) and k(cat)/K(m) values for the variants precluded any conclusions regarding a role for Asp-34 in the proton transfer step; however, the k(cat) of the D34A variant was chemically rescued by replacement of 2-(N-morpholino)propanesulfonic acid buffer with imidazole at pH 7.2, supporting a role for the conserved aspartate in the proton transfer step. The crystal structure of Cab also shows Arg-36 with two hydrogen bonds to Asp-34. Arg-36 variants had both k(cat) and k(cat)/K(m) values that were decreased at least 250-fold relative to those of the wild type, establishing an essential function for this residue. Imidazole was unable to rescue the k(cat) of the R36A variant; however, partial rescue of the kinetic parameter was obtained with guanidine-HCl indicating that the guanido group of this residue is important.  相似文献   

12.
Pan JJ  Yang LW  Liang PH 《Biochemistry》2000,39(45):13856-13861
Undecaprenyl pyrophosphate synthase (UPPs) catalyzes condensation of eight molecules of isopentenyl pyrophosphate with farnesyl pyrophosphate to yield C(55)-undecaprenyl pyrophosphate. We have mutated the aspartates and glutamates in the five conserved regions (I to V) of UPPs protein sequence to evaluate their effects on substrate binding and catalysis. The mutant enzymes including D26A, E73A, D150A, D190A, E198A, E213A, D218A, and D223A were expressed and purified to great homogeneity. Kinetic analyses of these mutant enzymes indicated that the substitution of D26 in region I with alanine resulted in a 10(3)-fold decrease of k(cat) value compared to wild-type UPPs. Its IPP K(m) value has only minor change. The mutagenesis of D150A has caused a much lower IPP affinity with IPP K(m) value 50-fold larger than that of wild-type UPPs but did not affect the FPP K(m) and the k(cat). The E213A mutant UPPs has a 70-fold increased IPP K(m) value and has a 100-fold decreased k(cat) value compared to wild-type. These results suggest that D26 of region I is critical for catalysis and D150 in region IV plays a significant role of IPP binding. The E213 residue in region V is also important in IPP binding as well as catalysis. Other mutant UPPs enzymes in this study have shown no significant change (<5-fold) of k(cat) with exception of E73A and D218A. Both enzymes have 10-fold lower k(cat) value relative to wild-type UPPs.  相似文献   

13.
Lietz EJ  Truher H  Kahn D  Hokenson MJ  Fink AL 《Biochemistry》2000,39(17):4971-4981
Lysine 73 is a conserved active-site residue in the class A beta-lactamases, as well as other members of the serine penicillin-sensitive enzyme family; its role in catalysis remains controversial and uncertain. Mutation of Lys73 to alanine in the beta-lactamase from Bacillus licheniformis resulted in a substantial reduction in both turnover rate (k(cat)) and catalytic efficiency (k(cat)/K(m)), and a very significant shift in pK(1) to higher pH in the bell-shaped pH-rate profiles (k(cat)/K(m)) for several penicillin and cephalosporin substrates. The increase in pK(1) is consistent with the removal of the positive ammonium group of the lysine from the proximity of Glu166, to which the acid limb has been ascribed. The alkaline limb of the k(cat)/K(m) vs profiles is not shifted appreciably, as might have been expected if this limb reflected the ionization of Lys73 in the wild-type enzyme. The k(cat)/K(m) at the pH optimum for the mutant was down about 200-fold for penicillins and around 10(4) for cephalosporins, compared to the wild-type, suggesting significant differences in the mechanisms for catalysis of penicillins compared to cephalosporins. Burst kinetics were observed with several substrates assayed with K73A beta-lactamase, indicating an underlying branched-pathway kinetic scheme, and rate-limiting deacylation. FTIR analysis was used to determine whether acylation or deacylation was rate-limiting. In general, acylation was the rate-limiting step for cephalosporin substrates, whereas deacylation was rate-limiting for penicillin substrates. The results indicate that Lys73 plays an important role in both the acylation and deacylation steps of the catalytic mechanism. The effects of this mutation (K73A) indicate that Lys73 does not function as a general base in the catalytic mechanism of beta-lactamase. The existence of bell-shaped pH-rate profiles for the K73A variant suggests that Lys73 is not directly responsible for either limb in such plots. It is likely that both Glu166 and Lys73 are important to each other in terms of maintaining the optimum electrostatic environment for fully efficient catalytic activity to occur.  相似文献   

14.
Hirano N  Haruki M  Morikawa M  Kanaya S 《Biochemistry》2000,39(43):13285-13294
A genetic method for isolating a mutant enzyme of ribonuclease HI (RNase HI) from Thermus thermophilus HB8 with enhanced activity at moderate temperatures was developed. T. thermophilus RNase HI has an ability to complement the RNase H-dependent temperature-sensitive (ts) growth phenotype of Escherichia coli MIC3001. However, this complementation ability was greatly reduced by replacing Asp(134), which is one of the active site residues, with His, probably due to a reduction in the catalytic activity. Random mutagenesis of the gene encoding the resultant D134H enzyme, followed by screening for second-site revertants, allowed us to isolate three single mutations (Ala(12) --> Ser, Lys(75) --> Met, and Ala(77) --> Pro) that restore the normal complementation ability to the D134H enzyme. These mutations were individually or simultaneously introduced into the wild-type enzyme, and the kinetic parameters of the resultant mutant enzymes for the hydrolysis of a DNA-RNA-DNA/DNA substrate were determined at 30 degrees C. Each mutation increased the k(cat)/K(m) value of the wild-type enzyme by 2.1-4.8-fold. The effects of the mutations on the enzymatic activity were roughly cumulative, and the combination of these three mutations increased the k(cat)/K(m) value of the wild-type enzyme by 40-fold (5.5-fold in k(cat)). Measurement of thermal stability of the mutant enzymes with circular dichroism spectroscopy in the presence of 1 M guanidine hydrochloride and 1 mM dithiothreitol showed that the T(m) value of the triple mutant enzyme, in which all three mutations were combined, was comparable to that of the wild-type enzyme (75.0 vs 77.4 degrees C). These results demonstrate that the activity of a thermophilic enzyme can be improved without a cost of protein stability.  相似文献   

15.
NAD(+)-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short-chain dehydrogenase/reductase (SDR) family, catalyzes the first step in the catabolic pathways of prostaglandins and lipoxins, and is believed to be the key enzyme responsible for the biological inactivation of these biologically potent eicosanoids. The enzyme utilizes NAD(+) specifically as a coenzyme. Potential amino acid residues involved in binding NAD(+) and facilitating enzyme catalysis have been partially identified. In this report, we propose that three more residues in 15-PGDH, Ile-17, Asn-91, and Val-186, are also involved in the interaction with NAD(+). Site-directed mutagenesis was used to examine their roles in binding NAD(+). Several mutants (I17A, I17V, I17L, I17E, I17K, N91A, N91D, N91K, V186A, V186I, V186D, and V186K) were prepared, expressed as glutathione S-transferase (GST) fusion enzymes in Escherichia coli, and purified by GSH-agarose affinity chromatography. Mutants I17E, I17K, N91L, N91K, and V186D were found to be inactive. Mutants N91A, N91D, V186A, and V186K exhibited comparable activities to the wild type enzyme. However, mutants I17A, I17V, I17L, and V186I had higher activity than the wild type. Especially, the activities of I17L and V186I were increased nearly 4- and 5-fold, respectively. The k(cat)/K(m) ratios of all active mutants for PGE(2) were similar to that of the wild type enzyme. However, the k(cat)/K(m) ratios of mutants I17A and N91A for NAD(+) were decreased 5- and 10-fold, respectively, whereas the k(cat)/K(m) ratios of mutants I17V, N91D, V186I, and V186K for NAD(+) were comparable to that of the wild type enzyme. The k(cat)/K(m) ratios of mutants I17L and V186A for NAD(+) were increased over nearly 2-fold. These results suggest that Ile-17, Asn-91, and Val-186 are involved in the interaction with NAD(+) and contribute to the full catalytic activity of 15-PGDH.  相似文献   

16.
Based on recent directed evolution of P450 2B1, six P450 2B11 mutants at three positions were created in an N-terminal modified construct termed P450 2B11dH and characterized for enzyme catalysis using five substrates. Mutant I209A demonstrated a 3.2-fold enhanced k(cat)/K(m) for 7-ethoxy-4-trifluoromethylcourmarin O-deethylation, largely due to a dramatic decrease in K(m) (0.72 microM vs. 18 microM). I209A also demonstrated enhanced selectivity for testosterone 16beta-hydroxylation over 16alpha-hydroxylation. In contrast, V183L showed a 4-fold increased k(cat) for 7-benzyloxyresorufin debenzylation and a 4.7-fold increased k(cat)/K(m) for testosterone 16alpha-hydroxylation. V183L also displayed a 1.7-fold higher k(cat)/K(m) than P450 2B11dH with the anti-cancer prodrugs cyclophosphamide and ifosfamide, resulting from a approximately 4-fold decrease in K(m). Introduction of the V183L mutation into full-length P450 2B11 did not enhance the k(cat)/K(m). Overall, the re-engineered P450 2B11dH enzymes exhibited enhanced catalytic efficiency with several substrates including the anti-cancer prodrugs.  相似文献   

17.
To elucidate the functions of the surface loops of VIIa, we prepared two mutants, VII-30 and VII-39. The VII-30 mutant had all of the residues in the 99 loop replaced with those of trypsin. In the VII-39 mutant, both the 99 and 170 loops were replaced with those of trypsin. The k(cat)/K(m) value for hydrolysis of the chromogenic peptidyl substrate S-2288 by VIIa-30 (103 mm(-)1s(-)1) was 3-fold higher than that of wild-type VIIa (30.3 mm(-)1 s(-)1) in the presence of soluble tissue factor (sTF). This enhancement was due to a decrease in the K(m) value but not to an increase in the k(cat) value. On the other hand, the k(cat)/K(m) value for S-2288 hydrolysis by VIIa-39 (17.9 mm(-)1 s(-)1) was 18-fold higher than that of wild-type (1.0 mm(-)1 s(-)1) in the absence of sTF, and the value was almost the same as that of wild-type measured in the presence of sTF. This enhancement was due to not only a decrease in the K(m) value but also to an increase in the k(cat) value. These results were in good agreement with their susceptibilities to a subsite 1-directed serine protease inhibitor. In our previous paper (Soejima, K., Mizuguchi, J., Yuguchi, M., Nakagaki, T., Higashi, S., and Iwanaga, S. (2001) J. Biol. Chem. 276, 17229-17235), the replacement of the 170 loop of VIIa with that of trypsin induced a 10-fold enhancement of the k(cat) value for S-2288 hydrolysis as compared with that of wild-type VIIa in the absence of sTF. These results suggested that the 99 and the 170 loop structures of VIIa independently affect the K(m) and k(cat) values, respectively. Furthermore, we studied the effect of mutations on proteolytic activity toward S-alkylated lysozyme as a macromolecular substrate and the activation of natural macromolecular substrate factor X.  相似文献   

18.
Specificity and kinetics of triose phosphate isomerase from chicken muscle   总被引:12,自引:6,他引:6  
The isolation of crystalline triose phosphate isomerase from chicken breast muscle is described. The values of k(cat.) and K(m) for the reaction in each direction were determined from experiments over wide substrate-concentration ranges, and the reactions were shown to obey simple Michaelis-Menten kinetics. With d-glyceraldehyde 3-phosphate as substrate, k(cat.) is 2.56x10(5)min(-1) and K(m) is 0.47mm; with dihydroxyacetone phosphate as substrate, k(cat.) is 2.59x10(4)min(-1) and K(m) is 0.97mm. The enzyme-catalysed exchange of the methyl hydrogen atoms of the ;virtual substrate' monohydroxyacetone phosphate with solvent (2)H(2)O or (3)H(2)O was shown. This exchange is about 10(4)-fold slower than the corresponding exchange of the C-3 hydrogen of dihydroxyacetone phosphate. The other deoxy substrate, 3-hydroxypropionaldehyde phosphate, was synthesized, but is too unstable in aqueous solution for analogous proton-exchange reactions to be studied.  相似文献   

19.
In the blood coagulation cascade, thrombin cleaves fibrinopeptides A and B from fibrinogen revealing sites for fibrin polymerization that lead to insoluble clot formation. Factor XIII stabilizes this clot by catalyzing the formation of intermolecular cross-links in the fibrin network. Thrombin activates the Factor XIII a(2) dimer by cleaving the Factor XIII activation peptide segment at the Arg(37)-Gly(38) peptide bond. Using a high performance liquid chromatography assay, the kinetic constants K(m), k(cat), and k(cat)/K(m) were determined for thrombin hydrolysis of fibrinogen Aalpha-(7-20), Factor XIII activation peptide-(28-41), and Factor XIII activation peptide-(28-41) with a Val(34) to Leu substitution. This Val to Leu mutation has been correlated with protection from myocardial infarction. In the absence of fibrin, the Factor XIII activation peptide-(28-41) exhibits a 10-fold lower k(cat)/K(m) value than fibrinogen Aalpha-(7-20). With the Factor XIII V34L mutation, decreases in K(m) and increases in k(cat) produce a 6-fold increase in k(cat)/K(m) relative to the wild-type Factor XIII sequence. A review of the x-ray crystal structures of known substrates and inhibitors of thrombin leads to a hypothesis that the new Leu generates a peptide with more extensive interactions with the surface of thrombin. As a result, the Factor XIII V34L is proposed to be susceptible to wasteful conversion of zymogen to activated enzyme. Premature depletion may provide cardioprotective effects.  相似文献   

20.
The deacetoxycephalosporin C synthase (DAOCS) from Streptomyces clavuligerus was engineered with the aim of enhancing the conversion of penicillin G into phenylacetyl-7-aminodeacetoxycephalosporanic acid, a precursor of 7-aminodeacetoxycephalosporanic acid, for industrial application. A single round of random mutagenesis followed by the screening of 5,500 clones identified three mutants, G79E, V275I, and C281Y, that showed a two- to sixfold increase in the k(cat)/K(m) ratio compared to the wild-type enzyme. Site-directed mutagenesis to modify residues surrounding the substrate resulted in three mutants, N304K, I305L, and I305M, with 6- to 14-fold-increased k(cat)/K(m) values. When mutants containing all possible combinations of these six sites were generated to optimize the ring expansion activity for penicillin G, the double mutant, YS67 (V275I, I305M), showed a significant 32-fold increase in the k(cat)/K(m) ratio and a 5-fold increase in relative activity for penicillin G, while the triple mutant, YS81 (V275I, C281Y, I305M), showed an even greater 13-fold increase in relative activity toward penicillin G. Our results demonstrate that this is a robust approach to the modification of DAOCS for an optimized DAOCS-penicillin G reaction.  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