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.
Comparisons of the structures of glutamate dehydrogenase (GluDH) and leucine dehydrogenase (LeuDH) have suggested that two substitutions, deep within the amino acid binding pockets of these homologous enzymes, from hydrophilic residues to hydrophobic ones are critical components of their differential substrate specificity. When one of these residues, K89, which hydrogen-bonds to the gamma-carboxyl group of the substrate l-glutamate in GluDH, was altered by site-directed mutagenesis to a leucine residue, the mutant enzyme showed increased substrate activity for methionine and norleucine but negligible activity with either glutamate or leucine. In order to understand the molecular basis of this shift in specificity we have determined the crystal structure of the K89L mutant of GluDH from Clostridium symbiosum. Analysis of the structure suggests that further subtle differences in the binding pocket prevent the mutant from using a branched hydrophobic substrate but permit the straight-chain amino acids to be used as substrates. The three-dimensional crystal structure of the GluDH from C. symbiosum has been previously determined in two distinct forms in the presence and absence of its substrate glutamate. A comparison of these two structures has revealed that the enzyme can adopt different conformations by flexing about the cleft between its two domains, providing a motion which is critical for orienting the partners involved in the hydride transfer reaction. It has previously been proposed that this conformational change is triggered by substrate binding. However, analysis of the K89L mutant shows that it adopts an almost identical conformation with that of the wild-type enzyme in the presence of substrate. Comparison of the mutant structure with both the wild-type open and closed forms has enabled us to separate conformational changes associated with substrate binding and domain motion and suggests that the domain closure may well be a property of the wild-type enzyme even in the absence of substrate.  相似文献   

2.
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   

3.
Horse liver alcohol dehydrogenase contains two tryptophan residues per subunit, Trp-15 on the surface of the catalytic domain and Trp-314 buried in the interface between the subunits of the dimer. We studied the contributions of the tryptophans to fluorescence and catalytic dynamics by substituting Trp-314 with a leucine residue and making two compensatory mutations that were required to obtain a stable protein, leading to the triple mutant M303F-L308I-W314L enzyme. The substitutions increased by two- to sixfold the turnover numbers for ethanol oxidation, acetaldehyde reduction, and the dissociation constants of the coenzymes. The rate of the exponential burst phase for the transient oxidation of ethanol increased slightly, but the rate of dissociation of the enzyme-NADH complex still limited turnover of ethanol, as for wild-type enzyme. The three substitutions at the dimer interface apparently activate the enzyme by allowing more rapid conformational changes that accompany coenzyme binding, probably due to movement of the loop containing residues 293 to 298. The emission spectrum of M303F-L308I-W314L enzyme, which contains Trp-15, was redshifted compared to wild-type enzyme. Time-resolved fluorescence measurements with the triple mutant show that the decay of Trp-15 is dominated by a approximately 7-ns component. In the mutant enzyme with Trp-15 substituted with phenylalanine, the decay of Trp-314 is dominated by a approximately 4-ns component. Solute quenching data for wild-type enzyme and the mutants show that only Trp-15 is exposed to iodide and acrylamide, whereas Trp-314 is inaccessible. The luminescence properties of the tryptophan residues in the mutated enzymes are consistent with conclusions from studies of the wild-type enzyme [M. R. Eftink, 1992, Adv. Biophys. Chem. 2, 81-114].  相似文献   

4.
A mutant tryptophan repressor (TrpR) protein containing the substitution of phenylalanine for leucine 75 has been isolated following a genetic screen for temperature-sensitive mutations. Two-dimensional (2D) 1H NMR spectra indicate an overall very similar fold for the purified mutant and wild-type proteins. Circular dichroism spectropolarimetry indicates an increased helix content relative to the wild-type protein, and a slightly higher urea denaturation midpoint for the mutant protein, although there is no difference in thermal stability. Fluorescence spectra indicate a more buried environment for one or both tryptophan residues in the mutant protein. The rate of proton-deuterium exchange-out for the resolved indole ring protons of the two tryptophan residues was quantified from NMR spectra of mutant and wild-type proteins and found to be approximately 50% faster in the wild-type protein. The mutant protein binds the corepressor l-tryptophan (l-Trp) approximately ten times more weakly than does the wild-type protein, but in l-Trp excess its DNA-binding affinity is only two to fivefold weaker. Taken together the results imply that, despite its conservative chemical character and surface location at the C terminus of helix one in the helix-turn-helix DNA recognition motif, this mutational change confers long-range effects on the dynamics of the protein's secondary and tertiary structure without substantially altering its fold, and with relatively minor effects on protein function.  相似文献   

5.
L1210/D3 mouse leukemia cells are resistant to 5, 10-dideazatetrahydrofolate due to expansion of cellular folate pools which block polyglutamation of the drug (Tse, A., and Moran, R. G. (1998) J. Biol. Chem. 273, 25944-25952). These cells were found to have two point mutations in the reduced folate carrier (RFC), resulting in a replacement of isoleucine 48 by phenylalanine and of tryptophan 105 by glycine. Each mutation contributes to the resistance phenotype. Genomic DNA from resistant cells contained both the wild-type and mutant alleles, but wild-type message was not detected. Folic acid was a much better substrate, and 5-formyltetrahydrofolate was a poorer substrate for transport in L1210/D3 cells relative to L1210 cells. Enhanced transport of folic acid was due to a marked, approximately 20-fold, decrease in the influx Km. Influx of methotrexate and 5,10-dideazatetrahydrofolate were minimally altered. Transfection of mutated rfc cDNA into RFC-null L1210/A cells produced the substrate specificity and 5, 10-dideazatetrahydrofolate resistance observed in the L1210/D3 line. Transfection of the mutant cDNA into wild-type cells also conferred resistance to 5,10-dideazatetrahydrofolate. We conclude that the I48F and W105G mutations in RFC caused resistance to 5, 10-dideazatetrahydrofolate, that the region of the RFC protein near these two positions defines the substrate-binding site, that the wild-type allele was silenced during the multistep development of resistance, and that this mutant phenotype represents a genetically dominant trait.  相似文献   

6.
Cyclodextrin glycosyltransferase (CGTase) is an industrially important enzyme that produces cyclodextrins (CD) from starch by intramolecular transglycosylation. CGTase consists of five globular domains labeled A through E. To better understand the role of domain E in CGTase catalysis, we have constructed several mutants of Bacillus macerans CGTase. Removing the entire E domain resulted in an inactive enzyme. Adding six amino acids between domains D and E caused a decrease in activity and thermostability. Replacing domain E with the similar starch-binding domain from Aspergillus awamori glucoamylase I caused a drastic decrease in activity, indicating the necessity of correct alignment of bound substrate. Substituting tyrosine residue 634 (Tyr634) with phenylalanine had very little effect on activity or thermostability. Substituting Tyr634 with glycine resulted in a 25% increase of specific cyclization and starch-hydrolyzing activities compared with that of the wild-type enzyme. The latter mutant was less thermostable. The results of this study indicate that domain E is important for the stability and integrity of B. macerans CGTase.  相似文献   

7.
The Rous sarcoma virus protease displays a high degree of specificity and catalyzes the cleavage of only a limited number of amino acid sequences. This specificity is governed by interactions between side chains of eight substrate amino acids and eight corresponding subsite pockets within the homodimeric enzyme. We have examined these complex interactions in order to learn how to introduce changes into the retroviral protease (PR) that direct it to cleave substrates. Mutant enzymes with altered substrate specificity and wild-type or greater catalytic rates have been constructed previously by substituting single key amino acids in each of the eight enzyme subsites with those residues found in structurally related positions of human immunodeficiency virus (HIV)-1 PR. These individual amino acid substitutions have now been combined into one enzyme, resulting in a highly active mutant Rous sarcoma virus (RSV) protease that displays many characteristics associated with the HIV-1 enzyme. The hybrid protease is capable of catalyzing the cleavage of a set of HIV-1 viral polyprotein substrates that are not recognized by the wild-type RSV enzyme. Additionally, the modified PR is inhibited completely by the HIV-1 PR-specific inhibitor KNI-272 at concentrations where wild-type RSV PR is unaffected. These results indicate that the major determinants that dictate RSV and HIV-1 PR substrate specificity have been identified. Since the viral protease is a homodimer, the rational design of enzymes with altered specificity also requires a thorough understanding of the importance of enzyme symmetry in substrate selection. We demonstrate here that the enzyme homodimer acts symmetrically in substrate selection with each enzyme subunit being capable of recognizing both halves of a peptide substrate equally.  相似文献   

8.
In thermolysin, tryptophan 115 seems to be at the S2 subsite. Trp-115 was replaced with tyrosine, phenylalanine, leucine, and valine during site-directed mutagenesis in order to evaluate the role of Trp-115 in the proteolytic activity of thermolysin. The mutant enzymes with Tyr-115 or Phe-115 had as much proteolytic activity as the wild-type enzyme, but the other two mutant enzymes had no activity. We found earlier that the substitution of Trp-115 with alanine, glutamic acid, lysine, and glutamine causes the enzyme to lose all activity, so an aromatic amino acid at position 115 seems to be essential for thermolysin.  相似文献   

9.
Dihydropyrimidine dehydrogenase catalyzes, in the rate-limiting step of the pyrimidine degradation pathway, the NADPH-dependent reduction of uracil and thymine to dihydrouracil and dihydrothymine, respectively. The porcine enzyme is a homodimeric iron-sulfur flavoprotein (2 x 111 kDa). C671, the residue postulated to be in the uracil binding site and to act as the catalytically essential acidic residue of the enzyme oxidative half-reaction, was replaced by an alanyl residue. The mutant enzyme was overproduced in Escherichia coli DH5alpha cells, purified to homogeneity, and characterized in comparison with the wild-type species. An extinction coefficient of 74 mM-1 cm-1 was determined at 450 nm for the wild-type and mutant enzymes. Chemical analyses of the flavin, iron, and acid-labile sulfur content of the enzyme subunits revealed similar stoichiometries for wild-type and C671A dihydropyrimidine dehydrogenases. One FAD and one FMN per enzyme subunit were found. Approximately 16 iron atoms and 16 acid-labile sulfur atoms were found per wild-type and mutant enzyme subunit. The C671A dihydropyrimidine dehydrogenase mutant exhibited approximately 1% of the activity of the wild-type enzyme, thus preventing its steady-state kinetic analysis. Therefore, the ability of the C671A mutant and, for comparison, of the wild-type enzyme species to interact with reaction substrates, products, or their analogues were studied by absorption spectroscopy. Both enzyme forms did not react with sulfite. The wild-type and mutant enzymes were very similar to each other with respect to the spectral changes induced by binding of the reaction product NADP+ or of its nonreducible analogue 3-aminopyridine dinucleotide phosphate. Uracil also induced qualitatively and quantitatively similar absorbance changes in the visible region of the absorbance spectrum of the two enzyme forms. However, the calculated Kd of the enzyme-uracil complex was significantly higher for the C671A mutant (9.1 +/- 0.7 microM) than for the wild-type dihydropyrimidine dehydrogenase (0.7 +/- 0.09 microM). In line with these observations, the two enzyme forms behaved in a similar way when titrated anaerobically with a NADPH solution. Addition of an up to 10-fold excess of NADPH to both dihydropyrimidine dehydrogenase forms led to absorbance changes consistent with reduction of approximately 0.5 flavin per subunit, with no indication of reduction of the enzyme iron-sulfur clusters. Absorbance changes consistent with reduction of both enzyme flavins were obtained by removing NADP+ with a NADPH-regenerating system. On the contrary, the two enzyme species differed significantly with respect to their reactivity with dihydrouracil. Addition of dihydrouracil to the wild-type enzyme species, under anaerobic conditions, led to absorbance changes that could be interpreted to result from both partial flavin reduction and the formation of a complex between the enzyme and (dihydro)uracil. In contrast, only spectral changes consistent with formation of a complex between the oxidized enzyme and dihydrouracil were observed when a C671A mutant enzyme solution was titrated with this compound. Furthermore, enzyme-monitored turnover experiments were carried out anaerobically in the presence of a limiting amount of NADPH and excess uracil with the two enzyme forms in a stopped-flow apparatus. These experiments directly demonstrated that the substitution of an alanyl residue for C671 in dihydropyrimidine dehydrogenase specifically prevents enzyme-catalyzed reduction of uracil. Finally, sequence analysis of dihydropyrimidine dehydrogenase revealed that it exhibits a modular structure; the N-terminal region, similar to the beta subunit of bacterial glutamate synthases, is proposed to be responsible for NADPH binding and oxidation with reduction of the FAD cofactor of dihydropyrimidine dehydrogenase. The central region, similar to the FMN subunit of dihydroorotate dehydrogenases, is likely to harbor the site o  相似文献   

10.
Xanthine dehydrogenase, a molybdenum, iron-sulfur flavoenzyme encoded in the fruit fly Drosophila melanogaster by the rosy gene, has been characterised both from the wild-type and mutant files. Enzyme assays, using a variety of different oxidising and reducing substrates were supplemented by limited molecular characterisation. Four rosy strains showed no detectable activity in any enzyme assay tried, whereas from four wild-type and three rosy mutant strains, those for the [E89K], [L127F] and [L157P]xanthine dehydrogenases (in all of which the mutation is in the iron-sulfur domain), the enzyme molecules, although present at different levels, had extremely similar or identical properties. This was confirmed by purification of one wild-type and one mutant enzyme. [E89K]xanthine dehydrogenase. These both had ultraviolet-visible absorption spectra similar to milk xanthine oxidase. Both were found to be quite stable molecules, showing very high catalytic-centre activities and with little tendency to become degraded by proteolysis or modified by conversion to oxidase or desulfo forms. In three further rosy strains, giving [G353D]xanthine dehydrogenase and [S357F]xanthine dehydrogenase mutated in the flavin domain, and [G1011E]xanthine dehydrogenase mutated in the molybdenum domain, enzyme activities were selectively diminished in certain assays. For the G353D and S357F mutant enzymes activities to NAD+ as oxidising substrate were diminished, to zero for the latter. In addition for [G353D]xanthine dehydrogenase, there was an increase in apparent Km values both for NAD+ and NADH. These findings indicate involvement of this part of the sequence in the NAD(+)-binding site. The G1011E mutation has a profound effect on the enzyme. As isolated and as present in crude extracts of the files, this xanthine dehydrogenase variant lacks activity to xanthine or pterin as reducing substrate, indicating an impairment of the functioning of its molybdenum centre. However, it retains full activity to NADH with dyes as oxidising substrate. Mild oxidation of the enzyme converts it, apparently irreversibly, to a form showing full activity to xanthine and pterin. The nature of the group that is oxidised is discussed in the light of redox potential data. It is proposed that the process involves oxidation of the pterin of the molybdenum cofactor from the tetrahydro to a dihydro oxidation state. This conclusion is fully consistent with recent information [Rom?o, M. J., Archer, M., Moura, I., Moura. J.J.G., LeGall, J., Engh, R., Schneider, M., Hof, P. & Huber, R. (1995) Science 270. 1170-1176) from X-ray crystallography on the structure of a closely related enzyme from Desulfovibrio gigas. It is proposed, that apparent irreversibility of the oxidative activating process for [G1011E]xanthine dehydrogenase, is due to conversion of its pterin to the tricyclic derivative detected by these workers. The data thus provide the strongest evidence available, that the oxidation state of the pterin can have a controlling influence on the activity of a molybdenum cofactor enzyme. Implications regarding pterin incorporation into xanthine dehydrogenase and in relation to other molybdenum enzymes are discussed.  相似文献   

11.
X-ray crystallography predicts hydrogen-bonding interactions between the side chains of Thr198 and two other amino acid residues, Glu194 (adjacent to the catalytic His195) and Ser318 (on the alpha-H helix which rearranges on substrate binding). In order to investigate the contribution of this conserved amino acid residue, Thr198, two mutants of Bacillus stearothermophilus lactate dehydrogenase were created (Val198 and Ile198). The steady-state kinetic parameters for both mutant enzymes were very similar with increased substrate Km and reduced kcat when compared with the wild-type enzyme. The mutation Val198 allowed non-productive binding of pyruvate to the unprotonated form of His195. Steady-state kinetic parameters determined for the Val198 mutant enzyme in high solvent viscosity suggested both an altered rate-limiting step in catalysis and implicated Thr198 in allosteric activation by the effector fructose 1,6-bisphosphate (Fru1,6P2). A shift in the Fru1,6P2 activation constant for the Val198 mutant enzyme suggested that Thr198 stabilises the catalytically competent (Fru1,6P2-activated) form of the enzyme by 6.6 kJ/mol. However, Thr198 was not important for maintaining the thermal stability of the Fru1,6P2-activated form. Equilibrium unfolding in guanidinium chloride indicated that Thr198 contributes 17.2 kJ/mol subunits towards the tertiary structural stability. The results emphasise the importance of the side chain-hydroxyl group of Thr198 which is required for (a) productive substrate binding, (b) allosteric activation and (c) protein conformational stability. The characteristics of the B. stearothermophilus lactate dehydrogenase mutations reported here were significantly different from those of the same mutations made in the corresponding position of the analogous enzyme Thermus flavus malate dehydrogenase [Nishiyama, M., Shimada, K., Horinouchi, S., & Beppu, T. (1991) J. Biol. Chem. 266, 14294-14299].  相似文献   

12.
The influence of the substrate P1' position on the specificity of two zinc matrix metalloproteases, membrane type-1 matrix metalloprotease (MT1-MMP) and stromelysin-3 (ST3), was evaluated by synthesizing a series of fluorogenic substrates of general formula dansyl-Pro-Leu-Ala-Xaa-Trp-Ala-Arg-NH2, where Xaa in the P1' position represents unusual amino acids containing either long arylalkyl or alkyl side chains. Our data demonstrate that both MT1-MMP and ST3 cleave substrates containing in their P1' position unusual amino acids with extremely long side chains more efficiently than the corresponding substrates with natural phenylalanine or leucine amino acids. In this series of substrates, the replacement of leucine by S-para-methoxybenzyl cysteine increased the kcat/Km ratio by a factor of 37 for MT1-MMP and 9 for ST3. The substrate with a S-para-methoxybenzyl cysteine residue in the P1' position displayed a kcat/Km value of 1.59 10(6) M-1 s-1 and 1.67 10(4) M-1 s-1, when assayed with MT1-MMP and ST3, respectively. This substrate is thus one of the most rapidly hydrolyzed substrates so far reported for matrixins, and is the first synthetic peptide efficiently cleaved by ST3. These unexpected results for these two matrixins suggest that extracellular proteins may be cleaved by matrixins at sites containing amino acids with unusual long side chains, like those generated in vivo by some post-translational modifications.  相似文献   

13.
The medium chain acyl-CoA dehydrogenase catalyzes the flavin-dependent oxidation of a variety of acyl-CoA thioesters with the transfer of reducing equivalents to electron-transferring flavoprotein. The binding of normal substrates profoundly suppresses the reactivity of the reduced enzyme toward molecular oxygen, whereas the oxidase reaction becomes significant using thioesters such as indolepropionyl-CoA (IP-CoA) and 4-(dimethylamino)-3-phenylpropionyl-CoA (DP-CoA). Steady-state and stopped-flow studies with IP-CoA led to a kinetic model of the oxidase reaction in which only the free reduced enzyme reacts with oxygen (Johnson, J. K., Kumar, N. R., and Srivastava, D. K. (1994) Biochemistry 33, 4738-4744). We have tested their proposal with IP-CoA and DP-CoA. The dependence of the oxidase reaction on oxygen concentration is biphasic with a major low affinity phase incompatible with a model predicting a simple Km for oxygen of 3 microM. If only free reduced enzyme reacts with oxygen, increasing IP-CoA would show strong substrate inhibition because it binds tightly to the reduced enzyme. Experimentally, IP-CoA shows simple saturation kinetics. The Glu376-Gln mutant of the medium chain dehydrogenase allows the oxygen reactivity of complexes of the reduced enzyme with IP-CoA and the corresponding product indoleacryloyl-CoA (IA-CoA) to be characterized without the subsequent redox equilibration that complicates analysis of the oxidase kinetics of the native enzyme. In sum, these data suggest that when bulky, nonphysiological substrates are employed, multiple reduced enzyme species react with molecular oxygen. The relatively high oxidase activity of the short chain acyl-CoA dehydrogenase from the obligate anaerobe Megasphaera elsdenii was studied by rapid reaction kinetics of wild-type and the Glu367-Gln mutant using butyryl-, crotonyl-, and 2-aza-butyryl-CoA thioesters. In marked contrast to those of the mammalian dehydrogenase, complexes of the reduced bacterial enzyme with these ligands react with molecular oxygen at rates similar to those of the free protein. Evolutionary and mechanistic aspects of the suppression of oxygen reactivity in the acyl-CoA dehydrogenases are discussed.  相似文献   

14.
Aminoacyl-tRNA synthetases activate amino acids with ATP to form aminoacyl adenylates as the essential intermediates for aminoacylation of their cognate tRNAs. The class I Escherichia coli cysteine tRNA synthetase contains an N-terminal nucleotide binding fold that provides the catalytic site of adenylate synthesis. The C-terminal domain of the cysteine enzyme is predominantly alpha-helical and contains a leucine heptad repeat motif. We show here that specific substitutions of leucines in the leucine heptad repeats reduced tRNA aminoacylation. In particular, substitution of Leu316 with phenylalanine reduced the catalytic efficiency of aminoacylation by 1000-fold. This deleterious effect was partially alleviated by a more conservative substitution of leucine with valine. Filter binding assays show that neither the phenylalanine nor the valine substitution at Leu316 had a major effect on the ability of the cysteine enzyme to bind tRNA(Cys). In contrast, pyrophosphate exchange assays show that both substitutions decreased the adenylate synthesis activity of the enzyme. Analysis of these results suggests that the primary defect of the valine substitution is executed at adenylate synthesis while that of the phenylalanine substitution is at both adenylate synthesis and the transition state of tRNA aminoacylation. Thus, although Leu316 is located in the C-terminal domain of the cysteine enzyme, it may modulate the capacity of the N-terminal domain for amino acid activation and tRNA aminoacylation through a domain-domain interaction.  相似文献   

15.
Phenylalanine ammonia-lyase (EC 4.3.1.5) from parsley is posttranslationally modified by dehydrating its Ser-202 to the catalytically essential dehydroalanine prosthetic group. The codon of Ser-202 was changed to those of alanine and threonine by site-directed mutagenesis. These mutants and the recombinant wild-type enzyme, after treatment with sodium borohydride, were virtually inactive with L-phenylalanine as substrate but catalyzed the deamination of L-4-nitrophenylalanine, which is also a substrate for the wild-type enzyme. Although the mutants reacted about 20 times slower with L-4-nitrophenylalanine than the wild-type enzyme, their Vmax for L-4-nitrophenylalanine was two orders of magnitude higher than for L-phenylalanine. In contrast to L-tyrosine, which was a poor substrate, DL-3-hydroxyphenylalanine (DL-m-tyrosine) was converted by phenylalanine ammonia-lyase at a rate comparable to that of L-phenylalanine. These results suggest a mechanism in which the crucial step is an electrophilic attack of the prosthetic group at position 2 or 6 of the phenyl group. In the resulting carbenium ion, the beta-HSi atom is activated in a similar way as it is in the nitro analogue. Subsequent elimination of ammonia, concomitant with restoration of both the aromatic ring and the prosthetic group, completes the catalytic cycle.  相似文献   

16.
L-Lactate dehydrogenase from Bifidobacterium longum shows homotropic activation by pyruvate as well as heterotropic activation by fructose 1,6-bisphosphate. Hybrid enzymes were produced from the wild-type subunit and a mutant subunit, whose substrate specificity was altered to that of malate dehydrogenase, and separated to analyze the substrate-induced homotropic activation mechanism. Oxamate, a competitive inhibitor of L-lactate dehydrogenase, was used to mimic the substrate-induced activation of the wild-type subunit as "a regulatory subunit." The malate dehydrogenase activity of the mutant subunit as "the catalytic subunit" of the hybrid enzymes was measured, and the activity of the mutant subunit was activated on the addition of oxamate. Thus, we directly observed the inter-subunit homotropic activation transmitted from the wild-type to the mutant subunit. Moreover, "isomeric" hybrid enzymes that have different structural subunit arrangements but identical subunit compositions showed identical kinetic natures. This indicates that the enzyme maintains its subunit symmetry during the allosteric transition.  相似文献   

17.
Mutation to the conserved Glu399 or Lys192 caused the rate-limiting step of human liver mitochondrial aldehyde dehydrogenase (ALDH2) to change from deacylation to hydride transfer (Sheikh, S., Ni, L., Hurley, T. D., and Weiner, H. (1997) J. Biol. Chem. 272, 18817-18822). Here we further investigated the role of these two NAD+-ribose-binding residues. The E399Q/K/H/D and K192Q mutants had lower dehydrogenase activity when compared with the native enzyme. No pre-steady state burst of NADH formation was found with the E399Q/K and K192Q enzymes when propionaldehyde was used as the substrate; furthermore, each mutant oxidized chloroacetaldehyde slower than propionaldehyde, and a primary isotope effect was observed for each mutant when [2H]acetaldehyde was used as a substrate. However, no isotope effect was observed for each mutant when alpha-[2H]benzaldehyde was the substrate. A pre-steady state burst of NADH formation was observed for the E399Q/K and K192Q mutants with benzaldehyde, and p-nitrobenzaldehyde was oxidized faster than benzaldehyde. Hence, when aromatic aldehydes were used as substrates, the rate-limiting step remained deacylation for all these mutants. The rate-limiting step remained deacylation for the E399H/D mutants when either aliphatic or aromatic aldehydes were used as substrates. The K192Q mutant displayed a change in substrate specificity, with aromatic aldehydes becoming better substrates than aliphatic aldehydes.  相似文献   

18.
We analyzed the role played by the conserved Gly154, a constituent of the P1 substrate-binding pocket of Bacillus subtilis subtilisin E, in the catalytic properties of the protease. Using an Escherichia coli expression system, the termination codon at position 154 in subtilisin E was first introduced to abolish the catalytic activity through truncation of the C-terminus from amino acid residues 154-275. We then attempted to obtain revertants with substitutions of various amino acids at position 154 by the polymerase chain reaction using a mixture of oligonucleotides. In addition to the Gly residue (wild-type), six amino acid substitutions (Ala, Arg, Leu, Phe, Pro and Thr) gave caseinolytic activity. When assayed with synthetic peptide substrates, most of the revertants showed a considerable decrease in specific activity and a P1 specificity similar to that of the wild-type enzyme. An Ala154 mutant purified from the periplasmic space in E. coli, however, resulted in an up to 2.3-fold preference for Val rather than Pro as a P2 substrate relative to the wild-type. Further, a significant 2-10-fold increase in the catalytic efficiency occurred in the Gly127Ala plus Gly154Ala combination variant, relative to the single Gly127Ala variant, without any change in the restricted specificity. The kinetic data and molecular modeling analysis demonstrate the important role of position 154 in the catalytic efficiency as well as in the substrate specificity of subtilisin E.  相似文献   

19.
Interactions of collagenases I and II (clostridiopeptidases) from Clostridium histolyticum with hexapeptide substrates in which some L-proline residues are replaced by their D-analogues, as well as with the tripeptide chloromethyl ketone Z-Gly-Pro-Gly-CH2Cl were studied. A role of stereochemistry of the amino acid residues in the substrate was established and differences between the collagenases, with regard to their specific requirements to substrates, were revealed. The tripeptide chloromethyl ketone is shown to be a specific collagenase inhibitor modifying at the substrate-binding site in the active centre of these enzymes, most likely lysine residues.  相似文献   

20.
The transsulfuration enzyme cystathionine gamma-synthase (CGS) catalyses the pyridoxal 5'-phosphate (PLP)-dependent gamma-replacement of O-succinyl-L-homoserine and L-cysteine, yielding L-cystathionine. The crystal structure of the Escherichia coli enzyme has been solved by molecular replacement with the known structure of cystathionine beta-lyase (CBL), and refined at 1.5 A resolution to a crystallographic R-factor of 20.0%. The enzyme crystallizes as an alpha4 tetramer with the subunits related by non-crystallographic 222 symmetry. The spatial fold of the subunits, with three functionally distinct domains and their quaternary arrangement, is similar to that of CBL. Previously proposed reaction mechanisms for CGS can be checked against the structural model, allowing interpretation of the catalytic and substrate-binding functions of individual active site residues. Enzyme-substrate models pinpoint specific residues responsible for the substrate specificity, in agreement with structural comparisons with CBL. Both steric and electrostatic designs of the active site seem to achieve proper substrate selection and productive orientation. Amino acid sequence and structural alignments of CGS and CBL suggest that differences in the substrate-binding characteristics are responsible for the different reaction chemistries. Because CGS catalyses the only known PLP-dependent replacement reaction at Cgamma of certain amino acids, the results will help in our understanding of the chemical versatility of PLP.  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