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.
Through comparison with the high-resolution structure of Clostridium symbiosum glutamate dehydrogenase, the different substrate specificities of the homologous enzymes phenylalanine dehydrogenase and leucine dehydrogenase were attributed to two residues, glycine 124 and leucine 307, in Bacillus sphaericus phenylalanine dehydrogenase, which are replaced with alanine and valine in leucine dehydrogenases. As predicted, making these substitutions in phenylalanine dehydrogenase decreased the specific activity towards aromatic substrates and enhanced the activity towards some aliphatic amino acids in standard assays with fixed concentrations of both substrates. This study did not, however, distinguish effects on affinity from those on maximum catalytic rate. A fuller kinetic characterization of the single- and double-mutant enzymes now reveals that the extent of the shift in specificity was underestimated in the earlier study. The maximum catalytic rates for aromatic substrates are reduced for all the mutants, but, in addition, the apparent Km values are higher for the single-mutant G124A and double-mutant G124A/L307V compared with the wild-type enzyme. Conversely, specificity constants (kcat/Km) for the nonpolar aliphatic amino acids and the corresponding 2-oxoacids for the mutants are all markedly higher than for the wild type, with up to a 40-fold increase for l-norvaline and a 100-fold increase for its 2-oxoacid in the double mutant. In some cases a favourable change in Km was found to outweigh a smaller negative change in kcat. These results emphasize the risk of misjudging the outcome of protein engineering experiments through too superficial an analysis. Overall, however, the success of the predictions from molecular modelling indicates the usefulness of this strategy for engineering new specificities, even in advance of more detailed 3D structural information.  相似文献   

2.
NAD+-dependent phenylalanine dehydrogenases were purified 1,500- and 1,600-fold, and crystallized from Sporosarcina ureae SCRC-R04 and Bacillus sphaericus SCRC-R79a, respectively. The purified enzymes were homogeneous as judged by disc gel electrophoresis. The enzyme from S. ureae has a molecular weight of 305,000, while that of B. sphaericus has a molecular weight of 340,000. Each is probably composed of eight subunits identical in molecular weight. The S. ureae enzyme showed a high substrate specificity in the oxidative deamination reaction acting on L-phenylalanine, while that of B. sphaericus acted on L-phenylalanine and L-tyrosine. The enzymes had lower substrate specificities in the reductive amination reaction acting on alpha-keto acids. The Sporosarcina enzyme acted on phenylpyruvate, alpha-ketocaproate, alpha-keto-gamma-methylthiobutyrate and rho-hydroxyphenylpyruvate. The Bacillus enzyme acted on rho-hydroxyphenylpyruvate, phenylpyruvate, and alpha-keto-gamma-methylthiobutyrate. The enzyme from B. sphaericus catalyzes The enzyme from B. sphaericus catalyzes the transfer of pro-S (B) hydrogen from NADH.  相似文献   

3.
The catabolism of branched chain amino acids, especially valine, appears to play an important role in furnishing building blocks for macrolide and polyether antibiotic biosyntheses. To determine the active site residues of ValDH, we previously cloned, partially characterized, and identified the active site (lysine) of Streptomyces albus ValDH. Here we report further characterization of S. albus ValDH. The molecular weight of S. albus ValDH was determined to be 38 kDa by SDS-PAGE and 67 kDa by gel filtration chromatography indicating that the enzyme is composed of two identical subunits. Optimal pHs were 10.5 and 8.0 for dehydrogenase activity with valine and for reductive amination activity with -ketoisovaleric acid, respectively. Several chemical reagents, which modify amino-acid side chains, inhibited the enzyme activity. To examine the role played by the residue for enzyme specificity, we constructed mutant ValDH by substituting alanine for glycine at position 124 by site-directed mutagenesis. This residue was chosen because it has been considered to be important for substrate discrimination by phenylalanine dehydrogenase (PheDH) and leucine dehydrogenase (LeuDH). The Ala-124–Gly mutant enzyme displayed lower activities toward aliphatic amino acids, but higher activities toward L-phenylalanine, L-tyrosine, and L-methionine compared to the wild type enzyme suggesting that Ala-124 is involved in substrate binding in S. albus ValDH.  相似文献   

4.
Wild-type phenylalanine dehydrogenase from Bacillus sphaericus, and three mutants N145A, N145V and N145L, are used with a coenzyme recycling system to synthesise L-phenylalanine and three non-natural amino acids (p-F-phenylalanine, p-MeO-phenylalanine and p-CF(3)-phenylalanine) on a millimole scale. A range of reaction conditions are investigated. The kinetic parameters of WT PheDH and N145A towards p-CF(3)-phenylpyruvate are compared, emphasising the value of protein engineering in creating improved biocatalysts.  相似文献   

5.
In eukaryotes, two isozymes (I and II) of methionine aminopeptidase (MetAP) catalyze the removal of the initiator methionine if the penultimate residue has a small radius of gyration (glycine, alanine, serine, threonine, proline, valine, and cysteine). Using site-directed mutagenesis, recombinant yeast MetAP I derivatives that are able to cleave N-terminal methionine from substrates that have larger penultimate residues have been expressed. A Met to Ala change at 329 (Met206 in Escherichia coli enzyme) produces an average catalytic efficiency 1.5-fold higher than the native enzyme on normal substrates and cleaves substrates containing penultimate asparagine, glutamine, isoleucine, leucine, methionine, and phenylalanine. Interestingly, the native enzyme also has significant activity with the asparagine peptide not previously identified as a substrate. Mutation of Gln356 (Gln233 in E. coli MetAP) to alanine results in a catalytic efficiency about one-third that of native with normal substrates but which can cleave methionine from substrates with penultimate histidine, asparagine, glutamine, leucine, methionine, phenylalanine, and tryptophan. Mutation of Ser195 to alanine had no effect on substrate specificity. None of the altered enzymes produced cleaved substrates with a fully charged residue (lysine, arginine, aspartic acid, or glutamic acid) or tyrosine in the penultimate position.  相似文献   

6.
N Okazaki  Y Hibino  Y Asano  M Ohmori  N Numao  K Kondo 《Gene》1988,63(2):337-341
The gene coding for phenylalanine dehydrogenase [PDH; L-phenylalanine: NAD+ oxidoreductase (deaminating); EC 1.4.1.-] from Bacillus sphaericus SCRC-79a was cloned onto plasmid pUC9, and the nucleotide sequence of the 2-kb DNA region of the insert was determined. A 1143-bp open reading frame consisting of 381 codons was identified as a pdh gene coding for PDH.  相似文献   

7.
The gene encoding the thermostable phenylalanine dehydrogenase [EC 1.4.1.-] of a thermophile, Thermoactinomyces intermedius, was cloned and its complete DNA sequence was determined. The phenylalanine dehydrogenase gene (pdh) consists of 1,098 nucleotides and encodes 366 amino acid residues corresponding to the subunit (Mr 41,000) of the hexameric enzyme. The amino acid sequence deduced from the nucleotide sequence of the pdh gene of T. intermedius was 56.0 and 42.1% homologous to those of the phenylalanine dehydrogenases of Bacillus sphaericus and Sporosarcina ureae, respectively. It shows 47.5% homology to that of the thermostable leucine dehydrogenase from B. stearothermophilus. The pdh gene was highly expressed in E. coli JM109, the amount of phenylalanine dehydrogenase produced amounting up to about 8.3% of that of the total soluble protein. We purified the enzyme to homogeneity from transformant cells in a day, with a 58% recovery.  相似文献   

8.
Recombinant rat insulysin was shown to cleave the internally quenched fluorogenic peptide 2-aminobenzyl-GGFLRKVGQ-ethylenediamine-2,4-dinitrophenol at the R-K bond, exhibiting a K(m) of 13 microm and a V(max) of 2.6 micromol min(-1) mg(-1). Derivatives of this peptide in which the P(2) leucine or the P(2)' valine were replaced with other residues were used to probe the subsite specificity of the enzyme. Varying the P(2) residue produced a 4-fold range in K(m) and a 7-fold range in k(cat). The nature of the P(2) residue had a significant effect on the site of cleavage. Leucine, isoleucine, valine, and aspartate produced cleavage at the R-K bond. Asparagine produced 36% cleavage at the N-R bond and 64% cleavage at the R-K bond, whereas with alanine or serine the A-R and S-R bonds were the major cleavage sites. With tyrosine, phenylalanine, methionine, or histidine representing the varied residue X, cleavages at F-X, X-R, and R-K were seen, whereas with tryptophan equal cleavage occurred at the F-W and W-R bonds. Variable P(2)' residues produce less of a change in both K(m) and k(cat) and have little influence on the cleavage site. Exceptions are phenylalanine, tyrosine, leucine, and isoleucine, which in addition to producing cleavage at the R-K bond, produce significant cleavage at the L-R bond. Alanine and tyrosine were unique in producing cleavage at the F-L bond. Taken together, these data suggest that insulysin specificity is directed toward the amino side of hydrophobic and basic residues and that the enzyme has an extended substrate binding site.  相似文献   

9.
Cloning and expression of the L-phenylalanine dehydrogenase (PheDH) gene from Bacillus sphaericus in B. subtilis was performed. It was ligated into the pHY300PLK shuttle vector and the resulting plasmid, pHYDH encoding polypeptide with molecular weight of 340 kDa, then transformed in B. subtilis ISW1214 and Escherichia coli JM109 competent cells for expression. Bacillus subtilis ISW1214/pHYDH only produced PheDH enzyme (4700 U/l). The recombinant PheDH was purified to near homogeneity as judged by SDS–polyacrylamide gel electrophoresis (M r 41000 Da) and the result was 40-fold with a yield of about 54%. Apparent K m values for L-phenylalanine (Phe), L-tyrosine and NAD+ were 0.24, 0.48 and 0.19 mM respectively. The optimum pH of the recombinant enzyme was 11 for the oxidative deamination, 10.2 for the reductive amination. The features of recombinant PheDH enzyme were comparable with the wild type PheDH protein.  相似文献   

10.
We describe a simple method for enzymatic synthesis of L and D amino acids from alpha-keto acids with Escherichia coli cells which express heterologous genes. L-amino acids were produced with thermostable L-amino acid dehydrogenase and formate dehydrogenase (FDH) from alpha-keto acids and ammonium formate with only an intracellular pool of NAD+ for the regeneration of NADH. We constructed plasmids containing, in addition to the FDH gene, the genes for amino acid dehydrogenases, including i.e., leucine dehydrogenase, alanine dehydrogenase, and phenylalanine dehydrogenase. L-Leucine, L-valine, L-norvaline, L-methionine, L-phenylalanine, and L-tyrosine were synthesized with the recombinant E. coli cells with high chemical yields (> 80%) and high optical yields (up to 100% enantiomeric excess). Stereospecific conversion of various alpha-keto acids to D amino acids was also examined with recombinant E. coli cells containing a plasmid coding for the four heterologous genes of the thermostable enzymes D-amino acid aminotransferase, alanine racemase, L-alanine dehydrogenase, and FDH. Optically pure D enantiomers of glutamate and leucine were obtained.  相似文献   

11.
In Escherichia coli, the active transport of phenylalanine is considered to be performed by two different systems, AroP and PheP. However, a low level of accumulation of phenylalanine was observed in an aromatic amino acid transporter-deficient E. coli strain (DeltaaroP DeltapheP Deltamtr Deltatna DeltatyrP). The uptake of phenylalanine by this strain was significantly inhibited in the presence of branched-chain amino acids. Genetic analysis and transport studies revealed that the LIV-I/LS system, which is a branched-chain amino acid transporter consisting of two periplasmic binding proteins, the LIV-binding protein (LIV-I system) and LS-binding protein (LS system), and membrane components, LivHMGF, is involved in phenylalanine accumulation in E. coli cells. The K(m) values for phenylalanine in the LIV-I and LS systems were determined to be 19 and 30 micro M, respectively. Competitive inhibition of phenylalanine uptake by isoleucine, leucine, and valine was observed for the LIV-I system and, surprisingly, also for the LS system, which has been assumed to be leucine specific on the basis of the results of binding studies with the purified LS-binding protein. We found that the LS system is capable of transporting isoleucine and valine with affinity comparable to that for leucine and that the LIV-I system is able to transport tyrosine with affinity lower than that seen with other substrates. The physiological importance of the LIV-I/LS system for phenylalanine accumulation was revealed in the growth of phenylalanine-auxotrophic E. coli strains under various conditions.  相似文献   

12.
The kinetics of isoleucine, leucine, and valine transport in Escherichia coli K-12 has been analyzed as a function of substrate concentration. Such analysis permits an operational definition of several transport systems having different affinities for their substrates. The identification of these transport systems was made possible by experiments on specific mutants whose isolation and characterization is described elsewhere. The transport process with highest affinity was called the "very-high-affinity"process. Isoleucine, leucine, and valine are substrates of this transport process and their apparent K(m) values are either 10(-8), 2 x 10(-8), or 10(-7) M, respectively. Methionine, threonine, and alanine inhibit this transport process, probably because they are also substrates. The very-high-affinity transport process is absent when bacteria are grown in the presence of methionine, and this is due to a specific repression. Methionine and alanine were also found to affect the pool size of isoleucine and valine. Another transport process is the "high-affinity" process. Isoleucine, leucine, and valine are substrates of this transport process, and their apparent K(m) value is 2 x 10(-6) M for all three. Methionine and alanine cause very little or no inhibition, whereas threonine appears to be a weak inhibitor. Several structural analogues of the branched-chain amino acids inhibit the very-high-affinity or the high-affinity transport process in a specific way, and this confirms their existence as two separate entities. Three different "low-affinity" transport processes, each specific for either isoleucine or leucine or valine, show apparent K(m) values of 0.5 x 10(-4) M. These transport processes show a very high substrate specificity since no inhibitor was found among other amino acids or among many branched-chain amino acid precursors or analogues tried. The evolutionary significance of the observed redundancy of transport systems is discussed.  相似文献   

13.
Cys-29 and Cys-251 of Streptomyces albus valine dehydrogenase (ValDH) were highly conserved in the corresponding region of NAD(P)(+)-dependent amino acid dehydroganase sequences. To ascertain the functional role of these cysteine residues in S. albus ValDH, site-directed mutagenesis was performed to change each of the two residues to serine. Kinetic analyses of the enzymes mutated at Cys-29 and Cys-251 revealed that these residues are involved in catalysis. We also constructed mutant ValDH by substituting valine for leucine at 305 by site-directed mutagenesis. This residue was chosen, because it has been proposed to be important for substrate discrimination by phenylalanine dehydrogenase (PheDH) and leucine dehydrogenase (LeuDH). Kinetic analysis of the V305L mutant enzyme revealed that it is involved in the substrate binding site. However it displayed less activity than the wild type enzyme toward all aliphatic and aromatic amino acids tested.  相似文献   

14.
Purified rat liver phenylalanine hydroxylase [L-phenylalanine:tetrahydropteridine:oxygen oxidoreductase (4-hydroxylating), EC 1.14.16.1] was immobilized with activated thiol-Sepharose 4B via disulfide bond formation, which is expected to immobilize the enzyme in its activated form through the SH modification. This immobilized enzyme was more stable against thermal denaturation than the free enzyme. When tetrahydrobiopterin was used as the natural cofactor, the K(m) value for phenylalanine was decreased and that for the cofactor was increased. Constant conversion from phenylalanine to tyrosine was demonstrated continuously for over 8 h at 25 degrees C.  相似文献   

15.
Phenylalanine pyruvate aminotransferase in rat liver was found in both the mitochondrial and supernatant fractions. Phenylalanine pyruvate aminotransferase was purified from rat liver mitochondria. The purified enzyme was specific for pyruvate, exhibiting no activity with 2-oxoglutarate as aminoacceptor, and utilized a wide range of amino acids as amino donors. Amino acids were effective in the following order of activity: L-phenylalanine > L-tyrosine > L-histidine > 3,4-dihydroxy-DL-phenylalanine. Very little activity was observed with L-tryptophan and 5-hydroxy-L-tryptophan. The apparent Km values for L-phenylalanine and L-histidine were 2.6 mM and 2.7 mM, respectively. The Km values for pyruvate were 5.0 mM and 1.5 mM with phenylalanine and histidine as amino donors, respectively. The pH optimum was near 9.0. Sucrose density gradient centrifugation gave a molecular weight of approximately 68,000. On the basis of subcellular distributions, substrate specificities, substrate inhibition, pH optima, polyacrylamide gel electrophoresis and some other properties, it was suggested that mitochondrial phenylalanine pyruvate aminotransferase was identical with mitochondrial histidine pyruvate aminotransferase.  相似文献   

16.
This study presents the partitioning and purification of recombinant Bacillus badius phenylalanine dehydrogenase (PheDH) in aqueous two-phase systems (ATPS) composed of polyethylene glycol 6000 (PEG-6000) and ammonium sulfate. A single-step operation of ATPS was developed for extraction and purification of recombinant PheDH from E. coli BL21 (DE3). The influence of system parameters including; PEG molecular weight and concentration, pH, (NH(4))(2)SO(4) concentration and NaCl salt addition on enzyme partitioning were investigated. The best optimal system for the partitioning and purification of PheDH was 8.5% (w/w) PEG-6000, 17.5% (w/w) (NH(4))(2)SO(4) and 13% (w/w) NaCl at pH 8.0. The partition coefficient, recovery, yield, purification factor and specific activity values were of 92.57, 141%, 95.85%, 474.3 and 10424.97 U/mg, respectively. Also the K(m) values for L-phenylalanine and NAD(+) in oxidative deamination were 0.020 and 0.13 mM, respectively. Our data suggested that this ATPS could be an economical and attractive technology for large-scale purification of recombinant PheDH.  相似文献   

17.
In earlier attempts to shift the substrate specificity of glutamate dehydrogenase (GDH) in favour of monocarboxylic amino-acid substrates, the active-site residues K89 and S380 were replaced by leucine and valine, respectively, which occupy corresponding positions in leucine dehydrogenase. In the GDH framework, however, the mutation S380V caused a steric clash. To avoid this, S380 has been replaced with alanine instead. The single mutant S380A and the combined double mutant K89L/S380A were satisfactorily overexpressed in soluble form and folded correctly as hexameric enzymes. Both were purified successfully by Remazol Red dye chromatography as routinely used for wild-type GDH. The S380A mutant shows much lower activity than wild-type GDH with glutamate. Activities towards monocarboxylic substrates were only marginally altered, and the pH profile of substrate specificity was not markedly altered. In the double mutant K89L/S380A, activity towards glutamate was undetectable. Activity towards L-methionine, L-norleucine and L-norvaline, however, was measurable at pH 7.0, 8.0 and 9.0, as for wild-type GDH. Ala163 is one of the residues that lines the binding pocket for the side chain of the amino-acid substrate. To explore its importance, the three mutants A163G, K89L/A163G and K89L/S380A/A163G were constructed. All three were abundantly overexpressed and showed chromatographic behaviour identical with that of wild-type GDH. With A163G, glutamate activity was lower at pH 7.0 and 8.0, but by contrast higher at pH 9.0 than with wild-type GDH. Activities towards five aliphatic amino acids were remarkably higher than those for the wild-type enzyme at pH 8.0 and 9.0. In addition, the mutant A163G used L-aspartate and L-leucine as substrates, neither of which gave any detectable activity with wild-type GDH. Compared with wild-type GDH, the A163 mutant showed lower catalytic efficiencies and higher K(m ) values for glutamate/2-oxoglutarate at pH 7.0, but a similar k(cat)/K(m) value and lower K(m) at pH 8.0, and a nearly 22-fold lower S(0.5) (substrate concentration giving half-saturation under conditions where Michaelis-Menten kinetics does not apply) at pH 9.0. Coupling the A163G mutation with the K89L mutation markedly enhanced activity (100-1000-fold) over that of the single mutant K89L towards monocarboxylic amino acids, especially L-norleucine and L-methionine. The triple mutant K89L/S380A/A163G retained a level of activity towards monocarboxylic amino acids similar to that of the double mutant K89L/A163G, but could no longer use glutamate as substrate. In terms of natural amino-acid substrates, the triple mutant represents effective conversion of a glutamate dehydrogenase into a methionine dehydrogenase. Kinetic parameters for the reductive amination reaction are also reported. At pH 7 the triple mutant and K89L/A163G show 5 to 10-fold increased catalytic efficiency, compared with K89L, towards the novel substrates. In the oxidative deamination reaction, it is not possible to estimate k(cat) and K(m) separately, but for reductive amination the additional mutations have no significant effect on k(cat) at pH 7, and the increase in catalytic efficiency is entirely attributable to the measured decrease in K(m). At pH 8 the enhancement of catalytic efficiency with the novel substrates was much more striking (e.g. for norleucine approximately 2000-fold compared with wild-type or the K89L mutant), but it was not established whether this is also exclusively due to more favourable Michaelis constants.  相似文献   

18.
Zhu Z  Sun D  Davidson VL 《Biochemistry》2000,39(37):11184-11186
Methylamine dehydrogenase (MADH) is a tryptophan tryptophylquinone (TTQ) dependent enzyme that catalyzes the oxidative deamination of primary amines. Amino acid residues of both the TTQ-bearing beta subunit and the noncatalytic alpha subunit line a substrate channel that leads from the protein surface to the enzyme active site. Phe55 of the alpha subunit is located at the opening of the active site. Conversion of alphaPhe55 to alanine dramatically alters the substrate preference of MADH. The K(m) for methylamine increases from 9 microM to 15 mM. The preferred substrates are now primary amines with chain lengths of at least seven carbons. The K(m) for 1, 10-diaminodecane is 11 microM, compared to 1.2 mM for wild-type MADH. Despite the large variation in K(m) values, k(cat) values are relatively unaffected by the mutation. Molecular modeling of substrates into the crystal structure of the enzyme active site and substrate channel provides an explanation for the dramatic changes in substrate specificity caused by this mutation of a single amino acid residue.  相似文献   

19.
Nitrilase from Rhodococcus rhodochrous ATCC 33278 hydrolyses both aliphatic and aromatic nitriles. Replacing Tyr-142 in the wild-type enzyme with the aromatic amino acid phenylalanine did not alter specificity for either substrate. However, the mutants containing non-polar aliphatic amino acids (alanine, valine and leucine) at position 142 were specific only for aromatic substrates such as benzonitrile, m-tolunitrile and 2-cyanopyridine, and not for aliphatic substrates. These results suggest that the hydrolysis of substrates probably involves the conjugated pi-electron system of the aromatic ring of substrate or Tyr-142 as an electron acceptor. Moreover, the mutants containing charged amino acids such as aspartate, glutamate, arginine and asparagine at position 142 displayed no activity towards any nitrile, possibly owing to the disruption of hydrophobic interactions with substrates. Thus aromaticity of substrate or amino acid at position 142 in R. rhodochrous nitrilase is required for enzyme activity.  相似文献   

20.
The Arabidopsis genome contains two genes predicted to code for bifunctional aspartate kinase-homoserine dehydrogenase enzymes (isoforms I and II). These two activities catalyze the first and the third steps toward the synthesis of the essential amino acids threonine, isoleucine, and methionine. We first characterized the kinetic and regulatory properties of the recombinant enzymes, showing that they mainly differ with respect to the inhibition of the homoserine dehydrogenase activity by threonine. A systematic search for other allosteric effectors allowed us to identify an additional inhibitor (leucine) and 5 activators (alanine, cysteine, isoleucine, serine, and valine) equally efficient on aspartate kinase I activity (4-fold activation). The six effectors of aspartate kinase I were all activators of aspartate kinase II activity (13-fold activation) and displayed a similar specificity for the enzyme. No synergy between different effectors could be observed. The activation, which resulted from a decrease in the Km values for the substrates, was detected using low substrates concentrations. Amino acid quantification revealed that alanine and threonine were much more abundant than the other effectors in Arabidopsis leaf chloroplasts. In vitro kinetics in the presence of physiological concentrations of the seven allosteric effectors confirmed that aspartate kinase I and II activities were highly sensitive to changes in alanine and threonine concentrations. Thus, physiological context rather than enzyme structure sets the specificity of the allosteric control. Stimulation by alanine may play the role of a feed forward activation of the aspartate-derived amino acid pathway in plant.  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