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.
Tyr225 in the active site of Escherichia coli aspartate aminotransferase (AspAT) was replaced by phenylalanine or arginine by site-directed mutagenesis. X-ray crystallographic analysis of Y225F AspAT showed that the benzene ring of Phe225 was situated at the same position as the phenol ring of Tyr225 in wild-type AspAT. The mutations resulted in a great decrease in the rate of the transamination reaction, suggesting that Tyr225 is important for efficient catalysis. The kinetic analysis of half-transamination reactions of Y225F AspAT with four substrates (aspartate, glutamate, oxalacetate, and 2-oxoglutarate) and some analogues (2-methylaspartate, succinate, and glutarate) revealed a considerable increase in the affinities for all these compounds. In contrast, affinity for the amino acid substrates was decreased by mutation to arginine, but affinities for the keto acid substrates and the two dicarboxylates (succinate and glutarate) were increased. The electrostatic interaction between O(3') of the coenzyme [pyridoxal 5'-phosphate (PLP)] and the residue at position 225 affected the pKa value of the Schiff base, which is formed between the epsilon-amino group of Lys258 and the aldehyde group of PLP; based on the spectrophotometric titration the pKa values were determined to be 6.8 for wild-type AspAT, 8.5 for Y225F AspAT, and 6.1 for Y225R AspAT in the absence of substrate. The absorption spectra of the three AspATs were almost identical in the acidic pH region, but the spectrum of Y225F AspAT differed from that of wild-type or Y225R AspAT in the alkaline pH region.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Asp222 is an invariant residue in all known sequences of aspartate aminotransferases from a variety of sources and is located within a distance of strong ionic interaction with N(1) of the coenzyme, pyridoxal 5'-phosphate (PLP), or pyridoxamine 5'-phosphate (PMP). This residue of Escherichia coli aspartate aminotransferase was replaced by Ala, Asn, or Glu by site-directed mutagenesis. The PLP form of the mutant enzyme D222E showed pH-dependent spectral changes with a pKa value of 6.44 for the protonation of the internal aldimine bond, slightly lower than that (6.7) for the wild-type enzyme. In contrast, the internal aldimine bond in the D222A or D222N enzyme did not titrate over the pH range 5.3-9.5, and a 430-nm band attributed to the protonated aldimine persisted even at high pH. The binding affinity of the D222A and D222N enzymes for PMP decreased by 3 orders of magnitude as compared to that of the wild-type enzyme. Pre-steady-state half-transamination reactions of all the mutant enzymes with substrates exhibited anomalous progress curves comprising multiphasic exponential processes, which were accounted for by postulating several kinetically different enzyme species for both the PLP and PMP forms of each mutant enzyme. While the replacement of Asp222 by Glu yielded fairly active enzyme species, the replacement by Ala and Asn resulted in 8600- and 20,000-fold decreases, respectively, in the catalytic efficiency (kmax/Kd value for the most active species of each mutant enzyme) in the reactions of the PLP form with aspartate. In contrast, the catalytic efficiency of the PMP form of the D222A or D222N enzyme with 2-oxoglutarate was still retained at a level as high as 2-10% of that of the wild-type enzyme. The presteady-state reactions of these two mutant enzymes with [2-2H]aspartate revealed a deuterium isotope effect (kH/kD = 6.0) greater than that [kH/kD = 2.2; Kuramitsu, S., Hiromi, K., Hayashi, H., Morino, Y., & Kagamiyama, H. (1990) Biochemistry 29, 5469-5476] for the wild-type enzyme. These findings indicate that the presence of a negatively charged residue at position 222 is particularly critical for the withdrawal of the alpha-proton of the amino acid substrate and accelerates this rate-determining step by about 5 kcal.mol-1. Thus it is concluded that Asp222 serves as a protein ligand tethering the coenzyme in a productive mode within the active site and stabilizes the protonated N(1) of the coenzyme to strengthen the electron-withdrawing capacity of the coenzyme.  相似文献   

3.
Site-directed mutagenesis of Tyr70 in the active site of Escherichia coli aspartate aminotransferase (AspAT) followed by kinetic studies has elucidated the roles of the hydroxyl group and benzene ring of Tyr70. X-ray crystallographic analysis showed that replacement of Tyr70 by Phe did not alter the active-site conformation of the enzyme. Comparison of the kinetic parameters of the four half-transamination reactions (the pyridoxal 5'-phosphate form of the enzyme with L-aspartate or L-glutamate and the pyridoxamine 5'-phosphate form with oxalacetate or 2-oxoglutarate) between the wild-type and [Tyr70----Phe]AspATs showed that the mutation increases the energy level of the transition state by 2 kcal.mol-1 for all the four substrates, suggesting some contribution of the hydroxyl group of Tyr70 to the transition state. When Phe70 was further replaced by Ser, the energy level of the transition state for L-glutamate or 2-oxoglutarate, but not for L-aspartate or oxalacetate, was further increased by 2-3 kcal.mol-1, suggesting that the presence of a benzene ring at position 70 is essential for recognizing the L-glutamate-2-oxoglutarate pair as substrates.  相似文献   

4.
P Carter  L Abrahmsén  J A Wells 《Biochemistry》1991,30(25):6142-6148
A mutant of the serine protease, subtilisin BPN', in which the catalytic His64 is replaced by Ala (H64A), is very specific for substrates containing a histidine, presumably by the substrate-bound histidine assisting in catalysis [Carter, P., & Wells, J.A. (1987) Science (Washington, D.C.) 237, 394-399]. Here we probe the catalytic mechanism of H64A subtilisin for cleaving His and non-His substrates. We show that the ratio of aminolysis to hydrolysis is the same for ester and amide substrates as catalyzed by the H64A subtilisin. This is consistent with formation of a common acyl-enzyme intermediate for H64A subtilisin, analogous to the mechanism of the wild-type enzyme. However, the catalytic efficiencies (kcat/KM) for amidase and esterase activities with His-containing substrates are reduced by 5000-fold and 14-fold, respectively, relative to wild-type subtilisin BPN, suggesting that acylation is more compromised than deacylation in the H64A mutant. High concentrations of imidazole are much less effective than His substrates in promoting hydrolysis by the H64A variant, suggesting that the His residue on the bound (not free) substrate is involved in catalysis. The reduction in catalytic efficiency kcat/KM for hydrolysis of the amide substrate upon replacement of the oxyanion stabilizing asparagine (N155G) is only 7-fold greater for wild-type than H64A subtilisin. In contrast, the reductions in kcat/KM upon replacement of the catalytic serine (S221A) or aspartate (D32A) are about 3000-fold greater for wild-type than H64A subtilisin, suggesting that the functional interactions between the Asp32 and Ser221 with the substrate histidine are more compromised in substrate-assisted catalysis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Prolyl 4-hydroxylase (EC 1.14.11.2), an alpha2beta2 tetramer, catalyzes the formation of 4-hydroxyproline in collagens. We converted 16 residues in the human alpha subunit individually to other amino acids, and expressed the mutant polypeptides together with the wild-type beta subunit in insect cells. Asp414Ala and Asp414Asn inactivated the enzyme completely, whereas Asp414Glu increased the K(m) for Fe2+ 15-fold and that for 2-oxoglutarate 5-fold. His412Glu, His483Glu and His483Arg inactivated the tetramer completely, as did Lys493Ala and Lys493His, whereas Lys493Arg increased the K(m) for 2-oxoglutarate 15-fold. His501Arg, His501Lys, His501Asn and His501Gln reduced the enzyme activity by 85-95%; all these mutations increased the K(m) for 2-oxoglutarate 2- to 3-fold and enhanced the rate of uncoupled decarboxylation of 2-oxoglutarate as a percentage of the rate of the complete reaction up to 12-fold. These and other data indicate that His412, Asp414 and His483 provide the three ligands required for the binding of Fe2+ to a catalytic site, while Lys493 provides the residue required for binding of the C-5 carboxyl group of 2-oxoglutarate. His501 is an additional critical residue at the catalytic site, probably being involved in both the binding of the C-1 carboxyl group of 2-oxoglutarate and the decarboxylation of this cosubstrate.  相似文献   

6.
Histidinol phosphate aminotransferase (HPAT) is a pyridoxal 5'-phosphate (PLP)-dependent aminotransferase classified into Subgroup I aminotransferase, in which aspartate aminotransferase (AspAT) is the prototype. In order to expand our knowledge on the reaction mechanism of Subgroup I aminotransferases, HPAT is an enzyme suitable for detailed mechanistic studies because of having low sequence identity with AspAT and a unique substrate recognition mode. Here we investigated the spectroscopic properties of HPAT and the effect of the C4-C4' strain of the PLP-Lys(214) Schiff base on regulating the Schiff base pK(a) in HPAT. Similar to AspAT, the PLP-form HPAT showed pH-dependent absorption spectral change with maxima at 340 nm at high pH and 420 nm at low pH, having a low pK(a) of 6.6. The pK(a) value of the methylamine-reconstituted K214A mutant enzyme was increased from 6.6 to 10.6. Mutation of Asn(157) to Ala increased the pK(a) to 9.2. Replacement of Arg(335) by Leu increased the pK(a) to 8.6. On the other hand, the pK(a) value of the N157A/R335L double mutant enzyme was 10.6. These data indicate that the strain of the Schiff base is the principal factor to decrease the pK(a) in HPAT and is crucial for the subsequent increase in the Schiff base pK(a) during catalysis, although the electrostatic effect of the arginine residue that binds the negatively charged group of the substrate is larger in HPAT than that in AspAT. Our findings also support the idea that the strain mechanism is common to Subgroup I aminotransferases.  相似文献   

7.
The four half-transamination reactions [the pyridoxal form of Escherichia coli aspartate aminotransferase (AspAT) with aspartate or glutamate and the pyridoxamine form of the enzyme with oxalacetate or 2-oxoglutarate] were followed in a stopped-flow spectrometer by monitoring the absorbance change at either 333 or 358 nm. The reaction progress curves in all cases gave fits to a monophasic exponential process. Kinetic analyses of these reactions showed that each half-reaction is composed of the following three processes: (1) the rapid binding of an amino acid substrate to the pyridoxal form of the enzyme; (2) the rapid binding of the corresponding keto acid to the pyridoxamine form of the enzyme; (3) the rate-determining interconversion between the two complexes. This mechanism was supported by the findings that the equilibrium constants for half- and overall-transamination reactions and the steady-state kinetic constants (Km and kcat) agreed well with the predicted values on the basis of the above mechanism using pre-steady-state kinetic parameters. The significant primary kinetic isotope effect observed in the reaction with deuterated amino acid suggests that the withdrawal of the alpha-proton of the substrates is rate determining. The pyridoxal form of E. coli AspAT reacted with a variety of amino acids as substrates. The Gibbs free energy difference between the transition state and the unbound state (unbound enzyme plus free substrate), as calculated from the pre-steady-state kinetic parameters, showed a linear relationship with the accessible surface area of amino acid substrate bearing an uncharged side chain.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Zheng R  Dam TK  Brewer CF  Blanchard JS 《Biochemistry》2004,43(22):7171-7178
Pantothenate synthetase (EC 6.3.2.1) catalyzes the formation of pantothenate from ATP, D-pantoate, and beta-alanine in bacteria, yeast, and plants. The three-dimensional structural determination of pantothenate synthetase from Mycobacterium tuberculosis has indicated specific roles for His44, His47, Asn69, Gln72, Lys160, and Gln164 residues in the binding of substrates and the pantoyl adenylate intermediate. To evaluate the functional roles of these strictly conserved residues, we constructed six Ala mutants and determined their catalytic properties. The substitution of alanine for H44, H47, N69, Q72, and K160 residues in M. tuberculosis pantothenate synthetase caused a greater than 1000-fold reduction in enzyme activity, while the Q164A mutant exhibited 50-fold less activity. The rate of the isolated adenylation reaction in single turnover studies was also reduced 40-1000-fold by the replacement of one of these six amino acids with alanine, suggesting that these residues are essential for the formation of the pantoyl adenylate intermediate. The rate of pantothenate formation from the adenylate and beta-alanine in the second half reaction could not be measured for the H44A, H47A, N69A, Q72A, and K160A mutants and was reduced 40-fold in the Q164A mutants. The activity of the K160C mutant enzyme was markedly enhanced by the alkylation of cysteine with bromoethylamine, further supporting the critical role of the K160 residue in pantoyl adenylate formation. Isothermal titration microcalorimetry analysis demonstrated that the substitution of either H47 or K160 for Ala resulted in a decreased affinity of the enzyme for ATP. These results indicate that the highly conserved His44, His47, Asn69, Gln72, Lys160 and residues are essential for the formation and stabilization of pantoyl adenylate intermediate in the pantothenate synthetase reaction.  相似文献   

9.
Substitution of a lysyl residue for Arg-386 of Escherichia coli aspartate aminotransferase resulted in an extensive decrease in Vmax values (0.8% with the aspartate-2-oxoglutarate pair and 0.2% with the glutamate-oxalacetate pair, compared with the corresponding values for the wild-type enzyme). Kinetic analysis of the four sets of half-reactions, the pyridoxal form of the enzyme with aspartate or glutamate and the pyridoxamine form with 2-oxoglutarate or oxalacetate, allowed us to define the independent effect of the mutation on the reactivity of each substrate. Decrease in the first order rate constant (kmax) was more pronounced in the reactions with five-carbon substrates (glutamate and 2-oxoglutarate) than in those with four-carbon substrates (aspartate and oxalacetate), while the increase in the apparent dissociation constant (Kd) was greater for four-carbon substrates than for five-carbon substrates. The decrease of overall catalytic efficiency as judged by the values, kmax/Kd, was more pronounced in the reactions with five-carbon substrates than in those with four-carbon substrates. Affinities for substrate analogs such as succinate, glutarate, 2-methylaspartate, and erythro-3-hydroxyaspartate, were also considerably decreased by the mutation of the enzyme. These findings indicate that the side chain of the lysyl residue, although it bears a positive charge similar to that of the arginyl residue, is not structurally adequate for the productive binding of a substrate during catalysis.  相似文献   

10.
Asn112 is located at the active site of thermolysin, 5-8 A from the catalytic Zn2+ and catalytic residues Glu143 and His231. When Asn112 was replaced with Ala, Asp, Glu, Lys, His, and Arg by site-directed mutagenesis, the mutant enzymes N112D and N112E, in which Asn112 is replaced with Asp and Glu, respectively, were secreted as an active form into Escherichia coli culture medium, while the other four were not. In the hydrolysis of a neutral substrate N-[3-(2-furyl)acryloyl]-Gly-L-Leu amide, the kcat/Km values of N112D and N112E exhibited bell-shaped pH-dependence, as did the wild-type thermolysin (WT). The acidic pKa of N112D was 5.7 +/- 0.1, higher by 0.4 +/- 0.2 units than that of WT, suggesting that the introduced negative charge suppressed the protonation of Glu143 or Zn2+-OH. In the hydrolysis of a negatively charged substrate, N-carbobenzoxy-l-Asp-l-Phe methyl ester (ZDFM), the pH-dependence of kcat/Km of the mutants decreased with increase in pH from 5.5 to 8.5, while that of WT was bell-shaped. This difference might be explained by the electrostatic repulsion between the introduced Asp/Glu and ZDFM, suggesting that introducing ionizing residues into the active site of thermolysin might be an effective means of modifying its pH-activity profile.  相似文献   

11.
Significant differences were found in the reaction rate, and the substrate and reaction specificities between the half reactions and the overall reactions catalyzed by L-lysine: 2-oxoglutarate 6-aminotransferase. The half reactions between an amino donor and the enzyme-bound pyridoxal 5'-phosphate, and also between an amino acceptor and the bound pyridoxamine 5'-phosphate followed first order reaction kinetics. The extrapolated first order rate constants and dissociation constants of the substrates were determined for the half reactions: lysine, 0.87 min-1 and 5.5 mM; glutamate, 1.1 min-1 and 10.5 mM; alanine, 0.66 min-1 and 6.6 mM; 6-aminohexanoate, 0.43 min-1 and 13.3 mM; and 2-oxoglutarate, 0.33 min-1 and 2.5 mM. As compared with the values reported for the overall reactions [Soda, K., Misono, H., & Yamamoto, T. (1968) Biochemistry 7, 4102-4109], the reactivity of the inherent substrates was lower by over 4 orders in the half reaction than that in the overall reaction, and the reactivity of alanine with the bound pyridoxal 5'-phosphate was reduced to 10% of that in the overall reaction. The substrate specificity in the half reaction was much lower than that in the overall reaction, which was re-examined in a reaction system containing the same concentration of the enzyme as that for the half reactions. Lysine 6-aminotransferase catalyzes the transfer of only the terminal amino group of lysine to 2-oxoglutarate in the overall reaction. However, in the half reaction, the 2-amino group as well as the terminal one was transferred to the bound pyridoxal 5'-phosphate. The ratio of reactivity of the 2-amino group to that of the 6-amino group was considerably influenced by the pH of the reaction mixture.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
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   

13.
The enantiomeric error frequency of aspartate aminotransferase (mitochondrial isoenzyme from chicken) was assessed by adding the enzyme in high concentration (0.89 mM) to a mixture of L-glutamate and 2-oxoglutarate (12 and 1.2 mM, respectively, at pH 7.5 and 25 degrees C). The substrates continuously undergo the transamination cycle under these conditions. Thereby, L-glutamate is progressively racemized, a 1:1 ratio of two enantiomers being reached within 240 h. The enantiomeric error frequency, i.e. the ratio of the rate of D-glutamate production and the rate of the transamination reaction with glutamate and 2-oxoglutarate as substrates, is 1.5 x 10(-7). D-Glutamate is also converted to a 1:1 racemic mixture. The racemizing activity of a mixture of free pyridoxal 5'-phosphate and pyridoxamine 5'-phosphate is about two orders of magnitude lower than that of aspartate aminotransferase. The error frequency of the enzyme in the case of the C4 substrate pair aspartate and oxalacetate is 3.4 x 10(-8), i.e. 4 times lower than that with the C5 substrate pair.  相似文献   

14.
Asn185 is an invariant residue in all known sequences of TPL and of closely related tryptophanase and it may be aligned with the Asn194 in aspartate aminotransferase. According to X-ray data, in the holoenzyme and in the Michaelis complex Asn185 does not interact with the cofactor pyridoxal 5'-phosphate, but in the external aldimine a conformational change occurs which is accompanied by formation of a hydrogen bond between Asn185 and the oxygen atom in position 3 of the cofactor. The substitution of Asn185 in TPL by alanine results in a mutant N185A TPL of moderate residual activity (2%) with respect to adequate substrates, L-tyrosine and 3-fluoro-L-tyrosine. The affinities of the mutant enzyme for various amino acid substrates and inhibitors, studied by both steady-state and rapid kinetic techniques, were lower than for the wild-type TPL. This effect mainly results from destabilization of the quinonoid intermediate, and it is therefore concluded that the hydrogen bond between Asn185 and the oxygen at the C-3 position of the cofactor is maintained in the quinonoid intermediate. The relative destabilization of the quinonoid intermediate and external aldimine leads to the formation of large amounts of gem-diamine in reactions of N185A TPL with 3-fluoro-L-tyrosine and L-phenylalanine. For the reaction with 3-fluoro-L-tyrosine it was first possible to determine kinetic parameters of gem-diamine formation by the stopped-flow method. For the reactions of N185A TPL with substrates bearing good leaving groups the observed values of k(cat) could be accounted for by taking into consideration two effects: the decrease in the quinonoid content under steady-state conditions and the increase in the quinonoid reactivity in a beta-elimination reaction. Both effects are due to destabilization of the quinonoid and they counterbalance each other. Multiple kinetic isotope effect studies on the reactions of N185A TPL with suitable substrates, L-tyrosine and 3-fluoro-L-tyrosine, show that the principal mechanism of catalysis, suggested previously for the wild-type enzyme, does not change. In the framework of this mechanism the observed considerable decrease in k(cat) values for reactions of N185A TPL with L-tyrosine and 3-fluoro-L-tyrosine may be ascribed to participation of Asn185 in additional stabilization of the keto quinonoid intermediate.  相似文献   

15.
VanXY(C), a bifunctional enzyme from VanC-phenotype Enterococcus gallinarum BM4174 that catalyses D,D-peptidase and D,D-carboxypeptidase activities, was purified as the native protein, as a maltose-binding protein fusion and with an N-terminal tag containing six histidine residues. The kinetic parameters of His(6)-VanXY(C) were measured for a variety of precursors of peptidoglycan synthesis involved in resistance: for D-Ala-D-Ala, the K(m) was 3.6 mm and k(cat), 2.5 s(-1); for UDP-MurNAc-L-Ala-D-Glu-L-Lys-DAla-D-Ala (UDP-MurNAc-pentapeptide[Ala]), K(m) was 18.8 mm and k(cat) 6.2 s(-1); for D-Ala-D-Ser, K(m) was 15.5 mm and k(cat) 0.35 s(-1). His(6)-VanXYC was inactive against the peptidoglycan precursor UDP-MurNAc-L-Ala-D-Glu-L-Lys-D-Ala-D-Ser (UDP-MurNAc-pentapeptide[Ser]). The rate of hydrolysis of the terminal D-Ala of UDP-MurNAc-pentapeptide[Ala] was inhibited 30% by 2 mm D-Ala-D-Ser or UDP-MurNAc-pentapeptide[Ser]. Therefore preferential hydrolysis of substrates terminating in D-Ala would occur during peptidoglycan synthesis in E. gallinarum BM4174, leaving precursors ending in D-Ser with a lower affinity for glycopeptides to be incorporated into peptidoglycan.Mutation of an aspartate residue (Asp59) of His-tagged VanXY(C) corresponding to Asp68 in VanX to Ser or Ala, resulted in a 50% increase and 73% decrease, respectively, of the specificity constant (k(cat)/K(m)) for D-Ala-D-Ala. This situation is in contrast to VanX in which mutation of Asp68-->Ala produced a greater than 200,000-fold decrease in the substrate specificity constant. This suggests that Asp59, unlike Asp68 in VanX, does not have a pivotal role in catalysis.  相似文献   

16.
The active site residue, His(15), in histidine-containing protein, HPr, can be replaced by aspartate and still act as a phosphoacceptor and phosphodonor with enzyme I and enzyme IIA(glucose), respectively. Other substitutions, including cysteine, glutamate, serine, threonine, and tyrosine, failed to show any activity. Enzyme I K(m) for His(15) --> Asp HPr is increased 10-fold and V(max) is decreased 1000-fold compared with wild type HPr. The phosphorylation of Asp(15) led to a spontaneous internal rearrangement involving the loss of the phosphoryl group and a water molecule, which was confirmed by mass spectrometry. The protein species formed had a higher pI than His(15) --> Asp HPr, which could arise from the formation of a succinimide or an isoimide. Hydrolysis of the isolated high pI form gave only aspartic acid at residue 15, and no isoaspartic acid was detected. This indicates that an isoimide rather than a succinimide is formed. In the absence of phosphorylation, no formation of the high pI form could be found, indicating that phosphorylation catalyzed the formation of the cyclization. The possible involvement of Asn(12) in an internal cyclization with Asp(15) was eliminated by the Asn(12) --> Ala mutation in His(15) --> AspHPr. Asn(12) substitutions of alanine, aspartate, serine, and threonine in wild type HPr indicated a general requirement for residues capable of forming a hydrogen bond with the Nepsilon(2) atom of His(15), but elimination of the hydrogen bond has only a 4-fold decrease in k(cat)/K(m).  相似文献   

17.
Aromatic amino acid aminotransferase is active toward both aromatic and dicarboxylic amino acids, and the mechanism for this dual substrate recognition has been an issue in the enzymology of this enzyme. Here we show that, in the reactions with aromatic and dicarboxylic ligands, the pK(a) of the Schiff base formed between the coenzyme pyridoxal 5'-phosphate and Lys258 or the substrate increases successively from 6.6 in the unliganded enzyme to approximately 8.8 in the Michaelis complex and to >10.5 in the external Schiff base complex. Mutations of Arg292 and Arg386 to Leu, which mimic neutralization of the positive charges of the two arginine residues by the ligand carboxylate groups, increased the Schiff base pK(a) by 0.1 and 0.7 unit, respectively. In contrast to these moderate effects of the Arg mutations, the cleavage of the Lys258 side chain of the Schiff base, which was brought about by preparing a mutant enzyme in which Lys258 was changed to Ala and the Schiff base was reconstituted with methylamine, produced the Schiff base pK(a) value of 10.2, that being 3.6 units higher than that of the wild-type enzyme. The observation indicates that the Schiff base pK(a) in the enzyme is lowered by the torsion around the C4-C4' axis of the Schiff base and suggests that the pK(a) is mainly controlled by changing the torsion angle during the course of catalysis. This mechanism, first observed for the reaction of aspartate aminotransferase with aspartate [Hayashi, H., Mizuguchi, H., and Kagamiyama, H. (1998) Biochemistry 37, 15076-15085], does not require the electrostatic contribution from the omega-carboxylate group of the substrate, and can explain why in aromatic amino acid aminotransferase the aromatic substrates can increase the Schiff base pK(a) during catalysis to the same extent as the dicarboxylic substrates. This is the first example in which the torsion pK(a) coupling of the pyridoxal 5'-phosphate Schiff base has been demonstrated in pyridoxal enzymes other than aspartate aminotransferase, and suggests the generality of the mechanism in the catalysis of aminotransferases related to aspartate aminotransferase.  相似文献   

18.
Measurement of the stereospecific release of the pro-S proton from C-4' of enzyme-bound pyridoxamine 5'-phosphate provides an experimental means to probe parts of the active site of aspartate aminotransferase independently of substrate turnover (Tobler, H. P., Christen, P., and Gehring, H. (1986) J. Biol. Chem. 261, 7105-7108). The release of pro-S 3H from enzyme-bound [3H]pyridoxamine 5'-phosphate is 30,000 times faster than from free coenzyme. Enzyme-bound [3H]pyridoxine 5'-phosphate is not detritiated suggesting an essential role of the 4'-amino group. Formation of the unproductive complex of the [3H]pyridoxamine 5'-phosphate-enzyme with aspartate or glutamate results in a 400-fold acceleration of 3H release. In contrast, addition of borohydride or cyanoborohydride immediately stops 3H release. Experiments with a fluorescent reporter group and with differential chemical modifications indicate that the activating effect of aspartate on the release of 3H is accompanied by a shift of the so-called open/closed conformational equilibrium of the enzyme (Kirsch, J.F., Eichele, G., Ford, G. C., Vincent, M.G., Jansonius, J.N., Gehring, H., and Christen, P. (1984) J. Mol. Biol. 174, 497-525) toward the closed conformation; the inhibiting effect of borohydride and cyanoborohydride appears to be accompanied by a shift toward the open conformation. Apparently, at least part of the catalytic apparatus of aspartate aminotransferase becomes fully operative only in the closed conformation of the enzyme.  相似文献   

19.
Human multidrug and toxic compound extrusion 1 (hMATE1) is an electroneutral H(+)/organic cation exchanger responsible for the final excretion step of structurally unrelated toxic organic cations in kidney and liver. To elucidate the molecular basis of the substrate recognition by hMATE1, we substituted the glutamate residues Glu273, Glu278, Glu300, and Glu389, which are conserved in the transmembrane regions, for alanine or aspartate and examined the transport activities of the resulting mutant proteins using tetraethylammonium (TEA) and cimetidine as substrates after expression in human embryonic kidney 293 (HEK-293) cells. All of these mutants except Glu273Ala were fully expressed and present in the plasma membrane of the HEK-293 cells. TEA transport activity in the mutant Glu278Ala was completely absent. Both Glu300Ala and Glu389Ala and all aspartate mutants exhibited significantly decreased activity. Glu273Asp showed higher affinity for cimetidine, whereas it has reduced affinity to TEA. Glu278Asp showed decreased affinity to cimetidine. Both Glu300Asp and Glu389Asp had lowered affinity to TEA, whereas the affinity of Glu389Asp to cimetidine was fourfold higher than that of the wild-type transporter with about a fourfold decrease in V(max) value. Both Glu273Asp and Glu300Asp had altered pH dependence for TEA uptake. These results suggest that all of these glutamate residues are involved in binding and/or transport of TEA and cimetidine but that their individual roles are different.  相似文献   

20.
We are probing the determinants of catalytic function and substrate specificity in serine proteases by kinetic and crystallographic characterization of genetically engineered site-directed mutants of rat trypsin. The role of the aspartyl residue at position 102, common to all members of the serine protease family, has been tested by substitution with asparagine. In the native enzyme, Asp102 accepts a hydrogen bond from the catalytic base His57, which facilitates the transfer of a proton from the enzyme nucleophile Ser195 to the substrate leaving group. At neutral pH, the mutant is four orders of magnitude less active than the naturally occurring enzyme, but its binding affinity for model substrates is virtually undiminished. Crystallographic analysis reveals that Asn102 donates a hydrogen bond to His57, forcing it to act as donor to Ser195. Below pH 6, His57 becomes statistically disordered. Presumably, the di-protonated population of histidyl side chains are unable to hydrogen bond to Asn102. Steric conflict may cause His57 to rotate away from the catalytic site. These results suggest that Asp102 not only provides inductive and orientation effects, but also stabilizes the productive tautomer of His57. Three experiments were carried out to alter the substrate specificity of trypsin. Glycine residues at positions 216 and 226 in the substrate-binding cavity were replaced by alanine residues in order to differentially affect lysine and arginine substrate binding. While the rate of catalysis by the mutant enzymes was reduced in the mutant enzymes, their substrate specificity was enhanced relative to trypsin. The increased specificity was caused by differential effects on the catalytic activity towards arginine and lysine substrates. The Gly----Ala substitution at 226 resulted in an altered conformation of the enzyme which is converted to an active trypsin-like conformation upon binding of a substrate analog. In a third experiment, Lys189, at the bottom of the specificity pocket, was replaced with an aspartate with the expectation that specificity of the enzyme might shift to aspartate. The mutant enzyme is not capable of cleaving at Arg and Lys or Asp, but shows an enhanced chymotrypsin-like specificity. Structural investigations of these mutants are in progress.  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