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 46 毫秒
1.
The roles of Asp(75), Asp(78), and Glu(83) of the (75)DPSDVARVE(83) element of Mycobacterium smegmatis GTP-dependent phosphoenolpyruvate (PEP) carboxykinase (GTP-PEPCK) were investigated. Asp(78) and Glu(83) are fully conserved in GTP-PEP-CKs. The human PEPCK crystal structure suggests that Asp(78) influences Tyr(220); Tyr(220) helps to position bound PEP, and Glu(83) interacts with Arg(81). Experimental data on other PEPCKs indicate that Arg(81) binds PEP, and the phosphate of PEP interacts with Mn(2+) of metal site 1 for catalysis. We found that D78A and E83A replacements severely reduced activity. E83A substitution raised the apparent K(m) value for Mn(2+) 170-fold. In contrast, Asp(75) is highly but not fully conserved; natural substitutions are Ala, Asn, Gln, or Ser. Such substitutions, when engineered, in M. smegmatis enzyme caused the following. 1) For oxaloacetate synthesis, V(max) decreased 1.4-4-fold. K(m) values for PEP and Mn(2+) increased 3-9- and 1.2-10-fold, respectively. K(m) values for GDP and bicarbonate changed little. 2) For PEP formation, V(max) increased 1.5-2.7-fold. K(m) values for oxaloacetate increased 2-2.8-fold. The substitutions did not change the secondary structure of protein significantly. The kinetic effects are rationalized as follows. In E83A the loss of Glu(83)-Arg(81) interaction affected Arg(81)-PEP association. D78A change altered the Tyr(220)-PEP interaction. These events perturbed PEP-Mn(2+) interaction and consequently affected catalysis severely. In contrast, substitutions at Asp(75), a site far from bound PEP, brought subtle effects, lowering oxaloacetate formation rate but enhancing PEP formation rate. It is likely that Asp(75) substitutions affected PEP-Mn(2+) interaction by changing the positions of Asp(78), Arg(81), and Glu(83), which translated to differential effects on two directions.  相似文献   

2.
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   

3.
Sobrado P  Fitzpatrick PF 《Biochemistry》2003,42(47):13833-13838
The flavoenzyme tryptophan 2-monooxygenase catalyzes the oxidation of tryptophan to indoleacetamide, carbon dioxide, and water. The enzyme is a homologue of l-amino acid oxidase. In the structure of l-amino acid oxidase complexed with aminobenzoate, Tyr372 hydrogen bonds with the carboxylate of the inhibitor in the active site. All 10 conserved tyrosine residues in tryptophan 2-monooxygenase were mutated to phenylalanine; steady state kinetic characterization of the purified proteins identified Tyr413 as the residue homologous to Tyr372 of l-amino acid oxidase. Y413F and Y413A tryptophan 2-monooxygenase were characterized more completely with tryptophan as the substrate to probe the contribution of this residue to catalysis. Mutation of Tyr413 to phenylalanine results in a decrease in the value of the first-order rate constant for reduction of 35-fold and a decrease in the rate constant for oxidation of 11-fold. Mutation to alanine decreases the rate constant for reduction by 200-fold and that for oxidation by 33-fold. Both mutations increase the K(d) value for tryptophan and the K(i) values for the competitive inhibitors indoleacetamide and indole pyruvate by 5-10-fold. Both mutations convert the enzyme to an oxidase, in that the products of the catalytic reactions of both are indolepyruvate and hydrogen peroxide. The V/K(trp)-pH profiles for the Tyr413 mutant enzymes no longer show the pK(a) value of 9.9 seen in that for the wild-type enzyme, allowing identification of Tyr413 as the active site residue in the wild-type enzyme which must be protonated for catalysis. Substitution of Tyr413 abolishes the formation of the long wavelength charge transfer species observed in the wild-type enzyme. The data are consistent with the main role of Tyr413 being to maintain the correct orientation of tryptophan for effective hydride transfer and imino acid decarboxylation.  相似文献   

4.
Kumar K  Walz FG 《Biochemistry》2001,40(12):3748-3757
Combinatorial random mutageneses involving either Asn43 with Asn44 (set 1) or Glu46 with an adjacent insertion (set 2) were undertaken to explore the functional perfection of the guanine recognition loop of ribonuclease T(1) (RNase T(1)). Four hundred unique recombinants were screened in each set for their ability to enhance enzyme catalysis of RNA cleavage. After a thorough selection procedure, only six variants were found that were either as active or more active than wild type which included substitutions of Asn43 by Gly, His, Leu, or Thr, an unplanned Tyr45Ser substitution and Glu46Pro with an adjacent Glu47 insertion. Asn43His-RNase T(1) has the same loop sequence as that for RNases Pb(1) and Fl(2). None of the most active mutants were single substitutions at Asn44 or double substitutions at Asn43 and Asn44. A total of 13 variants were purified, and these were subjected to kinetic analysis using RNA, GpC, and ApC as substrates. Modestly enhanced activities with GpC and RNA involved both k(cat) and K(M) effects. Mutants having low activity with GpC had proportionately even lower relative activity with RNA. Asn43Gly-RNase T(1) and all five of the purified mutants in set 2 exhibited similar values of k(cat)/K(M) for ApC which were the highest observed and about 10-fold that for wild type. The specificity ratio [(k(cat)/K(M))(GpC)/(k(cat)/K(M))(ApC)] varied over 30 000-fold including a 10-fold increase [Asn43His variant; mainly due to a low (k(cat)/K(M))(ApC)] and a 3000-fold decrease (Glu46Ser/(insert)Gly47 variant; mainly due to a low (k(cat)/K(M))(GpC)) as compared with wild type. It is interesting that k(cat) (GpC) for the Tyr45Ser variant was almost 4-fold greater than for wild type and that Pro46/(insert)Glu47 RNase T(1) is 70-fold more active than the permuted variant (insert)Pro47-RNase T(1) which has a conserved Glu46. In any event, the observation that only 6 out of 800 variants surveyed had wild-type activity supports the view that functional perfection of the guanine recognition loop of RNase T(1) has been achieved.  相似文献   

5.
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   

6.
The importance of the conserved Tyr352 and Asp380 residues of Bacillus stearothermophilus aminopeptidase II (AP-II) was investigated by site-directed mutagenesis. The wild-type and mutant enzymes were expressed in recombinant Escherichia coli M15 cells and the 45-kD proteins were purified from the cell-free extracts by Ni(2+)-NTA resin. The specific activity for Tyr352 and Asp380 replacements was decreased by more than 3.5-fold. Detailed analysis of the kinetic consequences in the mutant proteins revealed that the K (m) values were increased 1.9- to 2.6-fold with respect to wild-type enzyme. Catalytic efficiencies (k (cat)/K (m)) of mutant proteins were between 3.5- and 31-fold lower than the corresponding value of the wild-type enzyme. Tryptophan emission fluorescence and circular dichroism spectra were nearly identical for wild-type and mutant enzymes. These results indicate that residues Tyr352 and Asp380 are essential for the proper function of AP-II.  相似文献   

7.
Rishavy MA  Berkner KL 《Biochemistry》2008,47(37):9836-9846
Vitamin K-dependent (VKD) proteins become activated by the VKD carboxylase, which converts Glu's to carboxylated Glu's (Gla's) in their Gla domains. The carboxylase uses vitamin K epoxidation to drive Glu carboxylation, and the two half-reactions are coupled in 1:1 stoichiometry by an unknown mechanism. We now report the first identification of a residue, His160, required for coupling. A H160A mutant showed wild-type levels of epoxidation but substantially less carboxylation. Monitoring proton abstraction using a peptide with Glu tritiated at the gamma-carbon position revealed that poor coupling was due to impaired carbanion formation. H160A showed a 10-fold lower ratio of tritium release to vitamin K epoxidation than wild-type enzyme (i.e., 0.12 versus 1.14, respectively), which could fully account for the fold decrease in coupling efficiency. The Ala substitution in His160 did not affect the K m for vitamin K and caused only a 2-fold increase in the K m for Glu and 2-fold decrease in the activation of vitamin K epoxidation by Glu. The H160A K m for CO 2 was 5-fold higher than the wild-type enzyme. However, the k cat for H160A carboxylation was 8-9-fold lower than the wild-type enzyme with all three substrates (i.e., Glu, CO 2, and vitamin K), suggesting a catalytic role for His160 in carbanion formation. We propose that His160 facilitates the formation of the transition state for carbanion formation. His160 is highly conserved in metazoan VKD carboxylases but not in some bacterial orthologues (acquired by horizontal gene transfer), which has implications for how bacteria have adapted the carboxylase for novel functions.  相似文献   

8.
The contributions to substrate binding and catalysis of 13 amino acid residues of the Caenorhabditis elegans diadenosine tetraphosphate pyrophosphohydrolase (Ap(4)A hydrolase) predicted from the crystal structure of an enzyme-inhibitor complex have been investigated by site-directed mutagenesis. Sixteen glutathione S-transferase-Ap(4)A hydrolase fusion proteins were expressed and their k(cat) and K(m) values determined after removal of the glutathione S-transferase domain. As expected for a Nudix hydrolase, the wild type k(cat) of 23 s(-1) was reduced by 10(5)-, 10(3)-, and 30-fold, respectively, by replacement of the conserved P(4)-phosphate-binding catalytic residues Glu(56), Glu(52), and Glu(103) by Gln. K(m) values were not affected, indicating a lack of importance for substrate binding. In contrast, mutating His(31) to Val or Ala and Lys(83) to Met produced 10- and 16-fold increases in K(m) compared with the wild type value of 8.8 microm. These residues stabilize the P(1)-phosphate. H31V and H31A had a normal k(cat) but K83M showed a 37-fold reduction in k(cat). Lys(36) also stabilizes the P(1)-phosphate and a K36M mutant had a 10-fold reduced k(cat) but a relatively normal K(m). Thus both Lys(36) and Lys(83) may play a role in catalysis. The previously suggested roles of Tyr(27), His(38), Lys(79), and Lys(81) in stabilizing the P(2) and P(3)-phosphates were not confirmed by mutagenesis, indicating the absence of phosphate-specific binding contacts in this region. Also, mutating both Tyr(76) and Tyr(121), which clamp one substrate adenosine moiety between them in the crystal structure, to Ala only increased K(m) 4-fold. It is concluded that interactions with the P(1)- and P(4)-phosphates are minimum and sufficient requirements for substrate binding by this class of enzyme, indicating that it may have a much wider substrate range then previously believed.  相似文献   

9.
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   

10.
Attempts to alter the guanine specificity of ribonuclease T1 (RNase T1) by rational or random mutagenesis have failed so far. The RNase T1 variant RV (Lys41Glu, Tyr42Phe, Asn43Arg, Tyr45Trp, and Glu46Asn) designed by combination of a random and a rational mutagenesis approach, however, exhibits a stronger preference toward adenosine residues than wild-type RNase T1. Steady state kinetics of the cleavage reaction of the two dinucleoside phosphate substrates adenylyl-3',5'-cytidine and guanylyl-3',5'-cytidine revealed that the ApC/GpC ratio of the specificity coefficient (k(cat)/K(m)) was increased approximately 7250-fold compared to that of the wild-type. The crystal structure of the nucleotide-free RV variant has been refined in space group P6(1) to a crystallographic R-factor of 19.9% at 1.7 A resolution. The primary recognition site of the RV variant adopts a similar conformation as already known from crystal structures of RNase T1 not complexed to any nucleotide. Noteworthy is a high flexibility of Trp45 and Asn46 within the three individual molecules in the asymmetric unit. In addition to the kinetic studies, these data indicate the participation of Asn46 in the specific recognition of the base and therefore a specific binding of adenosine.  相似文献   

11.
Watling JR  Press MC  Quick WP 《Plant physiology》2000,123(3):1143-1152
We analyzed the impact of growth at either 350 (ambient) or 700 (elevated) microL L(-1) CO(2) on key elements of the C(4) pathway (photosynthesis, carbon isotope discrimination, and leaf anatomy) using the C(4) cereal sorghum (Sorghum bicolor L. Moench.). Gas-exchange analysis of the CO(2) response of photosynthesis indicated that both carboxylation efficiency and the CO(2) saturated rate of photosynthesis were lower in plants grown at elevated relative to ambient CO(2). This was accompanied by a 49% reduction in the phosphoenolpyruvate carboxylase content of leaves (area basis) in the elevated CO(2)-grown plants, but no change in Rubisco content. Despite the lower phosphoenolpyruvate carboxylase content, there was a 3-fold increase in C isotope discrimination in leaves of plants grown at elevated CO(2) and bundle sheath leakiness was estimated to be 24% and 33%, respectively, for the ambient and elevated CO(2)-grown plants. However, we could detect no difference in quantum yield. The ratio of quantum yield of CO(2) fixation to PSII efficiency was lower in plants grown at elevated CO(2), but only when leaf internal was below 50 microL L(-1). This suggests a reduction in the efficiency of the C(4) cycle when [CO(2)] is low, and also implies increased electron transport to acceptors other than CO(2). Analysis of leaf sections using a transmission electron microscope indicated a 2-fold decrease in the thickness of the bundle sheath cell walls in plants grown at elevated relative to ambient CO(2). These results suggest that significant acclimation to increased CO(2) concentrations occurs in sorghum.  相似文献   

12.
1. The properties of fructose diphosphatase from skeletal muscle of the Alaskan king-crab (Paralithodes camtschatica) were examined over the physiological temperature range of the animal. 2. King-crab muscle fructose diphosphatase is first activated by Na(+) and NH(4) (+) and is then partially inhibited by these cations at concentrations higher than 10mm at 0 degrees , 8 degrees and 15 degrees C. Enzyme activity is stimulated by K(+) at 0 degrees C, but is curtailed at 8 degrees C and 15 degrees C, an effect that could render rate independent of temperature. 3. Affinity for substrate increases with decreasing temperature; below the temperature of acclimatization, K(m) for fructose 1,6-diphosphate increases, resulting in a complex U-shaped temperature-K(m) curve. 4. King-crab muscle fructose diphosphatase is inhibited by low concentrations of AMP. As with enzymes of other poikilotherms, inhibition by AMP is sensitive to temperature; the enzyme is least sensitive to inhibition by AMP near the temperature of acclimatization. 5. The affinity of fructose diphosphatase for fructose 1,6-diphosphate is enhanced by phosphoenolpyruvate, and this activation is temperature-sensitive; 0.5mm-phosphoenolpyruvate causes a sevenfold decrease in K(m) for fructose 1,6-diphosphate at 15 degrees C but a 25-fold decrease at 0 degrees C. 6. Phosphoenolpyruvate appears to decrease the affinity of king-crab muscle fructose diphosphatase for AMP at low temperature, whereas at the higher temperature it appears to enhance inhibition by AMP. Phosphoenolpyruvate was not observed to cause a reversal of inhibition of fructose diphosphatase activity by AMP. The identification of phosphoenolpyruvate as an activator of a rate-limiting step in gluconeogenesis permits the suggestion of a coupling of the controlling mechanisms of several steps in the glycolytic and gluconeogenic chains. 7. These findings suggest mechanisms for the maintenance and regulation of control of fructose diphosphatase activity in king-crab skeletal muscle at low temperature and under conditions that favour concomitant activity of phosphofructokinase.  相似文献   

13.
In the crystal structure of chicken sulfite oxidase, the residue Tyr(322) (Tyr(343) in human sulfite oxidase) was found to directly interact with a bound sulfate molecule and was proposed to have an important role in mediating the substrate specificity and catalytic activity of this molybdoprotein. In order to understand the role of this residue in the catalytic mechanism of sulfite oxidase, steady-state and stopped-flow analyses were performed on wild-type and Y343F human sulfite oxidase over the pH range 6-10. In steady-state assays of Y343F sulfite oxidase using cytochrome c as the electron acceptor, k(cat) was somewhat impaired ( approximately 34% wild-type activity at pH 8.5), whereas the K(m)(sulfite) showed a 5-fold increase over wild type. In rapid kinetic assays of the reductive half-reaction of wild-type human sulfite oxidase, k(red)(heme) changed very little over the entire pH range, with a significant increase in K(d)(sulfite) at high pH. The k(red)(heme) of the Y343F variant was significantly impaired across the entire pH range, and unlike the wild-type protein, both k(red)(heme) and K(d)(sulfite) were dependent on pH, with a significant increase in both kinetic parameters at high pH. Additionally, reduction of the molybdenum center by sulfite was directly measured for the first time in rapid reaction assays using sulfite oxidase lacking the N-terminal heme-containing domain. Reduction of the molybdenum center was quite fast (k(red)(Mo) = 972 s(-1) at pH 8.65 for wild-type protein), indicating that this is not the rate-limiting step in the catalytic cycle. Reduction of the molybdenum center of the Y343F variant by sulfite was more significantly impaired at high pH than at low pH. These results demonstrate that the Tyr(343) residue is important for both substrate binding and oxidation of sulfite by sulfite oxidase.  相似文献   

14.
Epoxide hydrolase from Agrobacterium radiobacter catalyzes the hydrolysis of epoxides to their diols via an alkyl-enzyme intermediate. The recently solved X-ray structure of the enzyme shows that two tyrosine residues (Tyr152 and Tyr215) are positioned close to the nucleophile Asp107 in such a way that they can serve as proton donor in the alkylation reaction step. The role of these tyrosines, which are conserved in other epoxide hydrolases, was studied by site-directed mutagenesis. Mutation of Tyr215 to Phe and Ala and mutation of Tyr152 to Phe resulted in mutant enzymes of which the k(cat) values were only 2-4-fold lower than for wild-type enzyme, whereas the K(m) values for the (R)-enantiomers of styrene oxide and p-nitrostyrene oxide were 3 orders of magnitude higher than the K(m) values of wild-type enzyme, showing that the alkylation half-reaction is severely affected by the mutations. Pre-steady-state analysis of the conversion of (R)-styrene oxide by the Y215F and Y215A mutants showed that the 1000-fold elevated K(m) values were mainly caused by a 15-40-fold increase in K(S) and a 20-fold reduction in the rate of alkylation. The rates of hydrolysis of the alkyl-enzyme intermediates were not significantly affected by the mutations. The double mutant Y152F+Y215F showed only a low residual activity for (R)-styrene oxide, with a k(cat)/K(m) value that was 6 orders of magnitude lower than with wild-type enzyme and 3 orders of magnitude lower than with the single tyrosine mutants. This indicates that the effects of the mutations were cumulative. The side chain of Gln134 is positioned in the active site of the X-ray structure of epoxide hydrolase. Mutation of Gln134 to Ala resulted in an active enzyme with slightly altered steady-state kinetic parameters compared to wild-type enzyme, indicating that Gln134 is not essential for catalysis and that the side chain of Gln134 mimics bound substrate. Based upon this observation, the inhibitory potential of various unsubstituted amides was tested, resulting in the identification of phenylacetamide as a competitive inhibitor with an inhibition constant of 30 microM.  相似文献   

15.
Tyrosyl-tRNA synthetase (TyrRS) is able to catalyze the transfer of both l- and d-tyrosine to the 3' end of tRNA(Tyr). Activation of either stereoisomer by ATP results in formation of an enzyme-bound tyrosyl-adenylate intermediate and is accompanied by a blue shift in the intrinsic fluorescence of the protein. Single turnover kinetics for the aminoacylation of tRNA(Tyr) by D-tyrosine were monitored using stopped-flow fluorescence spectroscopy. Bacillus stearothermophilus tyrosyl-tRNA synthetase binds d-tyrosine with an 8.5-fold lower affinity than that of l-tyrosine (K (D-Tyr)(d) = 102 microm) and exhibits a 3-fold decrease in the forward rate constant for the activation reaction (k (D-Tyr)(3) = 13 s(-1)). Furthermore, as is the case for l-tyrosine, tyrosyl-tRNA synthetase exhibits "half-of-the-sites" reactivity with respect to the binding and activation of D-tyrosine. Surprisingly, pyrophosphate binds to the TyrRS.d-Tyr-AMP intermediate with a 14-fold higher affinity than it binds to the TyrRS.l-Tyr-AMP intermediate (K (PPi)(d) = 0.043 for TyrRS.d-Tyr-AMP.PP(i)). tRNA(Tyr) binds with a slightly (2.3-fold) lower affinity to the TyrRS.d-Tyr-AMP intermediate than it does to the TyrRS.l-Tyr-AMP intermediate. The observation that the K (Tyr)(d) and k(3) values are similar for l- and d-tyrosine suggests that their side chains bind to tyrosyl-tRNA synthetase in similar orientations and that at least one of the carboxylate oxygen atoms in d-tyrosine is properly positioned for attack on the alpha-phosphate of ATP.  相似文献   

16.
PEP Carboxykinase Exchange Reaction in Photosynthetic Bacteria   总被引:1,自引:1,他引:0  
This paper describes some new characteristics of the phosphoenolpyruvate carboxykinase CO(2)-oxaloacetate exchange reaction in purified preparations of Rhodospirillum rubrum. The enzymatic activity has been purified 169-fold. Nucleotide diphosphates substitute for nucleotide triphosphates in the exchange reaction. Nucleotide diphosphates will not support the synthesis of phosphoenolpyruvate from oxaloacetate. This reaction differs significantly from the CO(2)-oxaloacetate exchange reaction in higher plants and animals.  相似文献   

17.
The Ca(2+)/calmodulin-dependent endothelial cell myosin light chain kinase (MLCK) triggers actomyosin contraction essential for vascular barrier regulation and leukocyte diapedesis. Two high molecular weight MLCK splice variants, EC MLCK-1 and EC MLCK-2 (210-214 kDa), in human endothelium are identical except for a deleted single exon in MLCK-2 encoding a 69-amino acid stretch (amino acids 436-505) that contains potentially important consensus sites for phosphorylation by p60(Src) kinase (Lazar, V., and Garcia, J. G. (1999) Genomics 57, 256-267). We have now found that both recombinant EC MLCK splice variants exhibit comparable enzymatic activities but a 2-fold reduction of V(max), and a 2-fold increase in K(0.5 CaM) when compared with the SM MLCK isoform, whereas K(m) was similar in the three isoforms. However, only EC MLCK-1 is readily phosphorylated by purified p60(Src) in vitro, resulting in a 2- to 3-fold increase in EC MLCK-1 enzymatic activity (compared with EC MLCK-2 and SM MLCK). This increased activity of phospho-MLCK-1 was observed over a broad range of submaximal [Ca(2+)] levels with comparable EC(50) [Ca(2+)] for both phosphorylated and unphosphorylated EC MLCK-1. The sites of tyrosine phosphorylation catalyzed by p60(Src) are Tyr(464) and Tyr(471) within the 69-residue stretch deleted in the MLCK-2 splice variant. These results demonstrate for the first time that p60(Src)-mediated tyrosine phosphorylation represents an important mechanism for splice variant-specific regulation of nonmuscle MLCK and vascular cell function.  相似文献   

18.
Cho JH  Kim DH  Lee KJ  Kim DH  Choi KY 《Biochemistry》2001,40(34):10197-10203
We have investigated the function of Tyr248 using bovine wild-type CPA and its Y248F and Y248A mutants to find that the K(M) values were increased by 4.5-11-fold and the k(cat) values were reduced by 4.5-10.7-fold by the replacement of Tyr248 with Phe for the hydrolysis of hippuryl-L-Phe (HPA) and N-[3-(2-furyl)acryloyl]-Phe-Phe (FAPP), respectively. In the case of O-(trans-p-chlorocinnamoyl)-L-beta-phenyllactate (ClCPL), an ester substrate, the K(M) value was increased by 2.5-fold, and the k(cat) was reduced by 20-fold. The replacement of Tyr248 with Ala decreased the k(cat) values by about 18- and 237-fold for HPA and ClCPL, respectively, demonstrating that the aromatic ring of Tyr248 plays a critical role in the enzymic reaction. The increases of the K(M) values were only 6- and 5-fold for HPA and ClCPL, respectively. Thus, the present study indicates clearly that Tyr248 plays an important role not only in the binding of substrate but also in the enzymic hydrolysis. The kinetic results may be rationalized by the proposition that the phenolic hydroxyl of Tyr248 forms a hydrogen bond with the zinc-bound water molecule, causing further activation of the water molecule by reducing its pK(a) value. The pH dependency study of k(cat) values and the solvent isotope effects also support the proposition. A unified catalytic mechanism is proposed that can account for the different kinetic behavior observed in the CPA-catalyzed hydrolysis of peptide and ester substrates.  相似文献   

19.
This is the first report on a bacterial verterbrate-type GTP-dependent phosphoenolpyruvate carboxykinase (PCK). The pck gene of Mycobacterium smegmatis was cloned. The recombinant PCK was overexpressed in Escherichia coli in a soluble form and with high activity. The purified enzyme was found to be monomeric (72 kDa), thermophilic (optimum temperature, 70 degrees C), very stable upon storage at 4 degrees C, stimulated by thiol-containing reducing agents, and inhibited by oxalate and by alpha-ketoglutarate. The requirement for a divalent cation for activity was fulfilled best by Mn(2+) and Co(2+) and poorly by Mg(2+). At 37 degrees C, the highest V(m) value (32.5 units/mg) was recorded with Mn(2+) and in the presence of 37 mm dithiothreitol (DTT). The presence of Mg(2+) (2 mm) greatly lowered the apparent K(m) values for Mn(2+) (by 144-fold in the presence of DTT and by 9.4-fold in the absence of DTT) and Co(2+) (by 230-fold). In the absence of DTT but in the presence of Mg(2+) (2 mm) as the co-divalent cation, Co(2+) was 21-fold more efficient than Mn(2+). For producing oxaloacetate, the enzyme utilized both GDP and IDP; ADP served very poorly. The apparent K(m) values for phosphoenolpyruvate, GDP, and bicarbonate were >100, 66, and 8300 micrometer, respectively, whereas those for GTP and oxaloacetate (for the phosphoenolpyruvate formation activity) were 13 and 12 microm, respectively. Thus, this enzyme preferred the gluconeogenesis/glycerogenesis direction. This property fits the suggestion that in M. smegmatis, pyruvate carboxylase is not anaplerotic but rather gluconeogenic (Mukhopadhyay, B., and Purwantini, E. (2000) Biochim. Biophys. Acta. 1475, 191-206). Both in primary structure and kinetic properties, the mycobacterial PCK was very similar to its vertebrate-liver counterparts and thus could serve as a model for these enzymes; examples for several immediate targets are presented.  相似文献   

20.
Yoon HY  Lee EY  Cho SW 《Biochemistry》2002,41(21):6817-6823
The adenine binding domain of the ADP site within human glutamate dehydrogenase (GDH) was identified by cassette mutagenesis at the Tyr187 position. The wild type GDH was activated 3-fold by ADP at a concentration of 1 mM at pH 8.0, whereas no significant activation by ADP was observed with the Tyr187 mutant GDH regardless of the size, hydrophobicity, and ionization of the side chains. Studies of the steady-state velocity of the mutant enzymes revealed essentially unchanged apparent K(m) values for 2-oxoglutarate and NADH, but an approximately 4-fold decrease in the respective apparent V(max) values. The binding of ADP to the wild type or mutant GDH was further examined by photoaffinity labeling with [alpha-(32)P]8-azidoadenosine 5'-diphosphate (8N(3)ADP). 8N(3)ADP, without photolysis, mimicked the stimulatory properties of ADP on GDH activity. Saturation of photoinsertion with 8N(3)ADP occurred with apparent K(d) values near 25 microM for the wild type GDH, and the photoinsertion of [alpha-(32)P]8N(3)ADP was decreased best by ADP in comparison to other nucleotides. Unlike the wild type GDH, essentially no photoinsertion was detected for the Tyr187 mutant GDH in the presence or absence of 1 mM ADP. For the wild type GDH, photolabel-containing peptide generated by tryptic digestion was identified in the region containing the sequence EMSWIADTYASTIG, and the photolabeling of this peptide was prevented >95% by the presence of 1 mM ADP during photolysis, whereas no such a peptide was detected for the Tyr187 mutant GDH in the presence or absence of ADP. These results with cassette mutagenesis and photoaffinity labeling demonstrate selectivity of the photoprobe for the ADP binding site and suggest that the photolabeled peptide is within the ADP binding domain of the human GDH and that Tyr187 is responsible for the efficient base binding of ADP to human GDH.  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