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.
The Thermus thermophilus 3-isopropylmalate dehydrogenase (IPMDH) and Escherichia coli isocitrate dehydrogenase (ICDH) are two functionally and evolutionarily related enzymes with distinct substrate specificities. To understand the determinants of substrate specificities of the two proteins, the substrate and coenzyme in IPMDH were docked into their respective binding sites based on the published structure for apo IPMDH and its sequence and structural homology to ICDH. This modeling study suggests that (1) the substrate and coenzyme (NAD) binding modes of IPMDH are significantly different from those of ICDH, (2) the interactions between the substrates and coenzymes help explain the differences in substrate specificities of IPMDH and ICDH, and (3) binding of the substrate and coenzyme should induce a conformational change in the structure of IPMDH.  相似文献   

2.
Doyle SA  Fung SY  Koshland DE 《Biochemistry》2000,39(46):14348-14355
Despite the structural similarities between isocitrate and isopropylmalate, isocitrate dehydrogenase (IDH) exhibits a strong preference for its natural substrate. Using a combination of rational and random mutagenesis, we have engineered IDH to use isopropylmalate as a substrate. Rationally designed mutations were based on comparison of IDH to a similar enzyme, isopropylmalate dehydrogenase (IPMDH). A chimeric enzyme that replaced an active site loop-helix motif with IPMDH sequences exhibited no activity toward isopropylmalate, and site-directed mutants that replaced IDH residues with their IPMDH equivalents only showed small improvements in k(cat). Random mutants targeted the IDH active site at positions 113 (substituted with glutamate), 115, and 116 (both randomized) and were screened for activity toward isopropylmalate. Six mutants were identified that exhibited up to an 8-fold improvement in k(cat) and increased the apparent binding affinity by as much as a factor of 80. In addition to the S113E mutation, five other mutants contained substitutions at positions 115 and/or 116. Most small hydrophobic substitutions at position 116 improved activity, possibly by generating space to accommodate the isopropyl group of isopropylmalate; however, substitution with serine yielded the most improvement in k(cat). Only two substitutions were identified at position 115, which suggests a more specific role for the wild-type asparagine residue in the utilization of isopropylmalate. Since interactions between neighboring residues in this region greatly influenced the effects of each other in unexpected ways, structural solutions were best identified in combinations, as allowed by random mutagenesis.  相似文献   

3.
The crystal structure of the enzyme 3-isopropylmalate dehydrogenase (IPMDH) from Mycobacterium tuberculosis (LeuB, Mtb-IPMDH, Rv2995c) without substrate or co-factor was determined at 1.65 A resolution, which is the highest resolution reported for an IPMDH to date. The crystals contain two functional dimers in the asymmetric unit in an arrangement close to a tetramer of D2 symmetry. Despite the absence of a substrate or inhibitor bound to the protein, the structure of the monomer resembles the previously observed closed form of the enzyme more closely than the open form. A comparison with the substrate complex of IPMDH from Thiobacillus ferrooxidans and the co-factor complex of the Thermus thermophilus enzyme revealed a close relationship of the active-site architecture between the various bacterial enzymes. The inhibitor O-isobutenyl oxalylhydroxamate was found to bind to the active site of IPMDH in a mode similar to the substrate isopropylmalate.  相似文献   

4.
3-Isopropylmalate dehydrogenase (IPMDH) is a dimeric enzyme with a strongly hydrophobic core that is composed of residues from four alpha-helices. We replaced Glu253, which is found in the hydrophobic core and is part of the subunit interface of the Bacillus subtilis (Bs) IPMDH, with several other amino acids to probe. The thermostabilities of the mutants were assessed by measuring the residual enzymatic activities at 40 degrees C after heat treatment and by monitoring changes in ellipticity at 222 nm as the environmental temperature increased incrementally. The results of these studies indicate that, for residues with non-polar side chains, when positioned at residue 253, the thermostabilities of their corresponding mutants correlate positively with the relative hydrophobicities of the side chains. Relative activities of all mutants are lower than that of the wild-type enzyme. For two of the mutants, we directly show that the substitution at position 253 negatively affects Mn(2+) binding, which is required for catalysis. When a lysine is the position 253 residue, the protein dissociates. The results presented herein increase our understanding of the role played by the BsIPMDH dimer interface on the stability and activity of BsIPMDH.  相似文献   

5.
We have performed random mutagenesis coupled with selection to isolate mutant enzymes with high catalytic activities at low temperature from thermophilic 3-isopropylmalate dehydrogenase (IPMDH) originally isolated from Thermus thermophilus. Five cold-adapted mutant IPMDHs with single-amino-acid substitutions were obtained and analyzed. Kinetic analysis revealed that there are two types of cold-adapted mutant IPMDH: k(cat)-improved (improved in k(cat)) and K(m)-improved (improved in k(cat)/K(m)) types. To determine the mechanisms of cold adaptation of these mutants, thermodynamic parameters were estimated and compared with those of the Escherichia coli wild-type IPMDH. The Delta G(m) values for Michaelis intermediate formation of the k(cat)-improved-type enzymes were larger than that of the T. thermophilus wild-type IPMDH and similar to that of the E. coli wild-type IPMDH. The Delta G(m) values of K(m)-improved-type enzymes were smaller than that of the T. thermophilus wild-type IPMDH. Fitting of NAD(+) binding was improved in the K(m)-improved-type enzymes. The two types of cold-adapted mutants employed one of the two strategies of E. coli wild-type IPMDH: relative destabilization of the Michaelis complex in k(cat)-improved-type, and destabilization of the rate-limiting step in K(m)-improved type mutants. Some cold-adapted mutant IPMDHs retained thermostability similar to that of the T. thermophilus wild-type IPMDH.  相似文献   

6.
Redesigning of an enzyme for a new catalytic reaction and modified substrate specificity was exploited with 3-isopropylmalate dehydrogenase (IPMDH). Point-mutation on Gly-89, which is not in the catalytic site but near it, was done by changing it to Ala, Ser, Val, and Pro, and all the mutations changed the substrate specificity. The mutant enzymes showed higher catalytic efficiency (kcat/Km) than the native IPMDH when malate was used as a substrate instead of 3-isopropylmalate. More interestingly, an additional insertion of Gly between Gly-89 and Leu-90 significantly altered the substrate-specificity, although the overall catalytic activity was decreased. Particularly, this mutant turned out to efficiently accept D-lactic acid, which was not accepted as a substrate by wild-type IPMDH at all. These results demonstrate the opportunity for creating nove,enzymes by modification of amino acid residues that do not directly participate in catalysis, or by insertion of additional residues.  相似文献   

7.
A newly selected cold-adapted mutant 3-isopropylmalate dehydrogenase (IPMDH) from a random mutant library was a double mutant containing the mutations I11V and S92F that were found in cold-adapted mutant IPMDHs previously isolated. To elucidate the effect of each mutation on enzymatic activity, I11V and six multiple mutant IPMDHs were constructed and analyzed. All of the multiple mutant IPMDHs were found to be improved in catalytic activity at moderate temperatures by increasing the k(cat) with a simultaneous increase of K(m) for the coenzyme NAD(+). k(cat) was improved by a decrease in the activation enthalpy, DeltaH( not equal). The multiple mutants did not show large reduction in thermal stability, and one of them showed enhanced thermal stability. Mutation from I11 to V was revealed to have a stabilizing effect. Mutants showed increased thermal stability when the mutation I11V was combined. This indicates that it is possible to construct mutants with enhanced thermal stability by combining stabilizing mutation. No additivity was observed for the thermodynamic properties of catalytic reaction in the multiple mutant IPMDHs, implying that the structural changes induced by the mutations were interacting with each other. This indicates that careful and detailed tuning is required for enhancing activity in contrast to thermal stability.  相似文献   

8.
Random mutagenesis coupled with screening of the active enzyme at a low temperature was applied to isolate cold-adapted mutants of a thermophilic enzyme. Four mutant enzymes with enhanced specific activities (up to 4.1-fold at 40 degrees C) at a moderate temperature were isolated from randomly mutated Thermus thermophilus 3-isopropylmalate dehydrogenase. Kinetic analysis revealed two types of cold-adapted mutants, i.e. k(cat)-improved and K(m)-improved types. The k(cat)-improved mutants showed less temperature-dependent catalytic properties, resulting in improvement of k(cat) (up to 7.5-fold at 40 degrees C) at lower temperatures with increased K(m) values mainly for NAD. The K(m)-improved enzyme showed higher affinities toward the substrate and the coenzyme without significant change in k(cat) at the temperatures investigated (30-70 degrees C). In k(cat)-improved mutants, replacement of a residue was found near the binding pocket for the adenine portion of NAD. Two of the mutants retained thermal stability indistinguishable from the wild-type enzyme. Extreme thermal stability of the thermophilic enzyme is not necessarily decreased to improve the catalytic function at lower temperatures. The present strategy provides a powerful tool for obtaining active mutant enzymes at lower temperatures. The results also indicate that it is possible to obtain cold-adapted mutant enzymes with high thermal stability.  相似文献   

9.
We succeeded in further improvement of the stability of 3-isopropylmalate dehydrogenase (IPMDH) from an extreme thermophile, Thermus thermophilus, by a suppressor mutation method. We previously constructed a chimeric IPMDH consisting of portions of thermophile and mesophile enzymes. The chimeric enzyme is less thermostable than the thermophile enzyme. The gene encoding the chimeric enzyme was subjected to random mutagenesis and integrated into the genome of a leuB-deficient mutant of T. thermophilus. The transformants were screened at 76 degrees C in minimum medium, and three independent stabilized mutants were obtained. The leuB genes from these three mutants were cloned and analyzed. The sequence analyses revealed Ala-172-->Val substitution in all of the mutants. The thermal stability of the thermophile IPMDH was improved by introducing the amino acid substitution.  相似文献   

10.
The methionine chain-elongation pathway is required for aliphatic glucosinolate biosynthesis in plants and evolved from leucine biosynthesis. In Arabidopsis thaliana, three 3-isopropylmalate dehydrogenases (AtIPMDHs) play key roles in methionine chain-elongation for the synthesis of aliphatic glucosinolates (e.g. AtIPMDH1) and leucine (e.g. AtIPMDH2 and AtIPMDH3). Here we elucidate the molecular basis underlying the metabolic specialization of these enzymes. The 2.25 Å resolution crystal structure of AtIPMDH2 was solved to provide the first detailed molecular architecture of a plant IPMDH. Modeling of 3-isopropylmalate binding in the AtIPMDH2 active site and sequence comparisons of prokaryotic and eukaryotic IPMDH suggest that substitution of one active site residue may lead to altered substrate specificity and metabolic function. Site-directed mutagenesis of Phe-137 to a leucine in AtIPMDH1 (AtIPMDH1-F137L) reduced activity toward 3-(2′-methylthio)ethylmalate by 200-fold, but enhanced catalytic efficiency with 3-isopropylmalate to levels observed with AtIPMDH2 and AtIPMDH3. Conversely, the AtIPMDH2-L134F and AtIPMDH3-L133F mutants enhanced catalytic efficiency with 3-(2′-methylthio)ethylmalate ∼100-fold and reduced activity for 3-isopropylmalate. Furthermore, the altered in vivo glucosinolate profile of an Arabidopsis ipmdh1 T-DNA knock-out mutant could be restored to wild-type levels by constructs expressing AtIPMDH1, AtIPMDH2-L134F, or AtIPMDH3-L133F, but not by AtIPMDH1-F137L. These results indicate that a single amino acid substitution results in functional divergence of IPMDH in planta to affect substrate specificity and contributes to the evolution of specialized glucosinolate biosynthesis from the ancestral leucine pathway.  相似文献   

11.
The active site of sepiapterin reductase (SPR), which is a member of the NADP(H)-preferring short-chain dehydrogenase/reductase (SDR) family and acts as the terminal enzyme in the biosynthetic pathway of tetrahydrobiopterin cofactor (BH4), was investigated by truncation and site-directed mutagenesis. The truncation mutants showed that N-terminal and C-terminal residues contribute to bind coenzyme and substrate, respectively. The mutant rSPRA29V showed decreased activity; however, the A-X-L-L-S sequence, which has been reported as a putative pterin binding site, was estimated to preferably work as a component in the region for binding coenzyme rather than substrate. Site-directed mutants of rSPRS158D, rSPRY171V, and rSPRK175I showed low, but significant, activity having similar Km values and kcat/Km values less than 25%, for both sepiapterin and NADPH. Both amino acids Tyr-171 and Ser-158 are located within a similar distance to the carbonyl group of the substrate in the crystal structure of mouse SPR, and the double point mutant rSPRY171V+S158D was indicated to be inactive. These results showed that Ser-158, Tyr-171, and Lys-175 contributed to the catalytic activity of SPR, and both Tyr-171 and Ser-158 are simultaneously necessary on proton transfer to the carbonyl functional groups of substrate.  相似文献   

12.
Animal catechol O-methyltransferases and plant caffeoyl-coenzyme A O-methyltransferases share about 20% sequence identity and display common structural features. The crystallographic structure of rat liver catechol O-methyltransferase was used as a template to construct a homology model for tobacco caffeoyl-coenzyme A O-methyltransferase. Integrating substrate specificity data, the three-dimensional model identified several amino acid residues putatively involved in substrate binding. These residues were mutated by a polymerase chain reaction method and wild-type and mutant enzymes were each expressed in Escherichia coli and purified. Substitution of Arg-220 with Thr resulted in the total loss of enzyme activity, thus indicating that Arg-220 is involved in the electrostatic interaction with the coenzyme A moiety of the substrate. Changes of Asp-58 to Ala and Gln-61 to Ser were shown to increase K(m) values for caffeoyl coenzyme A and to decrease catalytic activity. Deletions of two amino acid sequences specific for plant enzymes abolished activity. The secondary structures of the mutants, as measured by circular dichroism, were essentially unperturbed as compared with the wild type. Similar changes in circular dichroism spectra were observed after addition of caffeoyl coenzyme A to the wild-type enzyme and the substitution mutants but not in the case of deletion mutants, thus revealing the importance of these sequences in substrate-enzyme interactions.  相似文献   

13.
To characterize the thermal stability of 3-isopropylmalate dehydrogenase (IPMDH) from an extreme thermophile, Thermus thermophilus, urea-induced unfolding of the enzyme and of its mesophilic counterpart from Escherichia coli was investigated at various temperatures. The unfolding curves were analyzed with a three-state model for E.coli IPMDH and with a two-state model for T.thermophilus IPMDH, to obtain the free energy change DeltaG degrees of each unfolding process. Other thermodynamic parameters, enthalpy change DeltaH, entropy change DeltaS and heat capacity change DeltaC(p), were derived from the temperature dependence of DeltaG degrees. The main feature of the thermophilic enzyme was its lower dependence of DeltaG degrees on temperature resulting from a low DeltaC(p). The thermophilic IPMDH had a significantly lower DeltaC(p), 1.73 kcal/mol.K, than that of E.coli IPMDH (20.7 kcal/mol.K). The low DeltaC(p) of T.thermophilus IPMDH could not be predicted from its change in solvent-accessible surface area DeltaASA. The results suggested that there is a large structural difference between the unfolded state of T.thermophilus and that of E.coli IPMDH. Another responsible factor for the higher thermal stability of T.thermophilus IPMDH was the increase in the most stable temperature T(s). The DeltaG degrees maximum of T.thermophilus IPMDH was much smaller than that of E.coli IPMDH. The present results clearly demonstrated that a higher melting temperature T(m) is not necessarily accompanied by a higher DeltaG degrees maximum.  相似文献   

14.
Watanabe K  Yamagishi A 《FEBS letters》2006,580(16):3867-3871
Previously, we showed that mutants of Thermus thermophilus 3-isopropylmalate dehydrogenase (IPMDH) each containing a residue (ancestral residue) that had been predicted to exist in a postulated common ancestor protein often have greater thermal stabilities than does the contemporary wild-type enzyme. In this study, the combined effects of multiple ancestral residues were analyzed. Two mutants, containing multiple mutations, Sup3mut (Val181Thr/Pro324Thr/Ala335Glu) and Sup4mut (Leu134Asn/Val181Thr/Pro324Thr/Ala335Glu) were constructed and show greater thermal stabilities than the wild-type and single-point mutant IPMDHs do. Most of the mutants have similar or improved catalytic efficiencies at 70 degrees C when compared with the wild-type IPMDH.  相似文献   

15.
1. A rapid method for estimating the activity of the first enzyme of lysine biosynthesis in yeasts (acetyl-coenzyme A: 2-ketoglutarate C-acetyl transferase, EC 4.1.3.21) is described. 2. In the wild type strain, the fixation of one substrate, S-acetyl coenzyme A, shows sigmoidal saturation kinetics. The initial rate experiments indicate that the reaction obeys an ordered mechanism, 2-ketoglutaric acid binding before S-acetyl coenzyme A. 3. The activity is completely inhibited in vitro by lysine and by some lysine analogs, which all show cooperative binding and have an heterotropic effect on 2-ketoglutaric binding sites. A second class of affectors is found, including 2-aminoadipic acid, pipecolic acid and dipicolinic acid, which all affect the cooperativity of S-acetyl coenzyme A binding sites. 4. Two types of mutations which modify these inhibition patterns without affecting the catalytic activity are described. One results in a desensitization towards lysine and lysine analogs only. The other entirely abolishes the susceptibility towards the second type of inhibitors, without affecting the susceptibility to lysine. 5. No variations of the specific activity could be detected in the wild type strain at all; mutants showing an increased or a reduced activity were isolated. 6. Our results do not support the existence of isoenzymes at the level of homocitrate synthetase in this yeast.  相似文献   

16.
Gox2253 from Gluconobacter oxydans belongs to the short‐chain dehydrogenases/reductases family, and catalyzes the reduction of heptanal, octanal, nonanal, and decanal with NADPH. To develop a robust working platform to engineer novel G. oxydans oxidoreductases with designed coenzyme preference, we adopted a structure based rational design strategy using computational predictions that considers the number of hydrogen bonds formed between enzyme and docked coenzyme. We report the crystal structure of Gox2253 at 2.6 Å resolution, ternary models of Gox2253 mutants in complex with NADH/short‐chain aldehydes, and propose a structural mechanism of substrate selection. Molecular dynamics simulation shows that hydrogen bonds could form between 2′‐hydroxyl group in the adenosine moiety of NADH and the side chain of Gox2253 mutant after arginine at position 42 is replaced with tyrosine or lysine. Consistent with the molecular dynamics prediction, Gox2253‐R42Y/K mutants can use both NADH and NADPH as a coenzyme. Hence, the strategies here could provide a practical platform to engineer coenzyme selectivity for any given oxidoreductase and could serve as an additional consideration to engineer substrate‐binding pockets. Proteins 2014; 82:2925–2935. © 2014 Wiley Periodicals, Inc.  相似文献   

17.
Circular dichroism spectra and circular dichroism difference spectra, generated when porcine heart mitochondrial and supernatant malate dehydrogenase bind coenzymes or when enzyme dihydroincotinamide nucleotide binary complexes bind substrate analogs, are presented. No significant changes are observed in protein chromophores in the 200- to 240-nm spectral range indicating that there is apparently little or no perturbation of the alpha helix or peptide backbone when binary or ternary complexes are formed. Quite different spectral perturbances occur in the two enzymes with reduced coenzyme binding as well as with substrate-analog binding by enzyme-reduced coenzyme binding. Comparison of spectral perturbations in both enzymes with oxidized or reduced coenzyme binding suggests that the dihydronicotinamide moiety of the coenzyme interacts with or perturbs indirectly the environment of aromatic amino acid residues. Reduced coenzyme binding apparently perturbs tyrosine residues in both mitochondrial malate dehydrogenase and lactic dehydrogenase. Reduced coenzyme binding perturbs tyrosine and tryptophan residues in supernatant malate dehydrogenase. The number of reduced coenzyme binding sites was determined to be two per 70,000 daltons in the mitochondrial enzyme, and the reduced coenzyme dissociation constants, determined through the change in ellipticity at 260 nm, with dihydronicotinamide adenine dinucleotide binding, were found to be good agreement with published values (Holbrook, J. J., and Wolfe, R. G. (1972) Biochemistry 11, 2499-2502) obtained through fluorescence-binding studies and indicate no apparent extra coenzyme binding sites. When D-malate forms a ternary complex with malate dehydrogenase-reduced coenzyme complexes, perturbation of both adenine and dihydronicotinamide chromophores is evident. L-Malate binding, however, apparently produces only a perturbation of the adenine chromophore in such complexes. Since the coenzyme has been found to bind in an open conformation on the surface of the enzyme and the substrate analogs bind at or very near the dihydronicotinamide moiety binding site, protein conformational changes are implicated during ternary complex formation with D-malate which can effect the adenine chromophore at some distance from the substrate binding site.  相似文献   

18.
Xia L  Ballou DP  Marsh EN 《Biochemistry》2004,43(11):3238-3245
Arginine-100 is involved in recognizing the gamma carboxylate of the substrate in glutamate mutase. To investigate its role in substrate binding and catalysis, this residue was mutated to lysine, tyrosine, and methionine. The effect of these mutations was to reduce k(cat) by 120-320-fold and to increase K(m(apparent)) for glutamate by 13-22-fold; K(m(apparent)) for adenosylcobalamin is little changed by these mutations. Even at saturating substrate concentrations, no cob(II)alamin could be detected in the UV-visible spectra of the Arg100Tyr and Arg100Met mutants. However, in the Arg100Lys mutant cob(II)alamin accumulated to concentrations similar to wild-type enzyme, which allowed the pre-steady-state kinetics of adenosylcobalamin homolysis to be investigated by stopped-flow spectroscopy. It was found that homolysis of the coenzyme is slower by an order of magnitude, compared with wild-type enzyme. Furthermore, glutamate binding is significantly weakened, so much so that the reaction exhibits second-order kinetics over the range of substrate concentrations used. The Arg100Lys mutant does not exhibit the very large deuterium isotope effects that are observed for homolysis of the coenzyme when the wild-type enzyme is reacted with deuterated substrates; this suggests that homolysis is slowed relative to hydrogen abstraction by this mutation.  相似文献   

19.
The kinetic mechanism and the substrate specificity of liver alcohol dehydrogenase are changed when 3-benzoylpyridine-adenine dinucleotide is used as coenzyme. Only primary alcohols are substrates of the enzyme and with ethanol the mechanism becomes rapid-equilibrium random bi-bi. According to model building experiments on a graphic display, the benzoyl group partially enters the substrate binding site, whereas the essential interactions between coenzyme and enzyme are preserved. This restraint on the substrate binding site provides a molecular explanation for the observed dependence between coenzyme and substrate chemical structures.  相似文献   

20.
Ancestral amino acid residues were inferred for 3-isopropylmalate dehydrogenase (IPMDH), and were introduced into the enzyme of an extreme thermophile, Sulfolobus sp. strain 7. The thermostability of the mutant enzymes was compared with that of the wild type enzyme. At least five of the seven mutants tested showed higher thermal stability than the wild type IPMDH. The results are compatible with the hyperthermophilic universal ancestor hypothesis. The results also provide a new method for designing thermostable enzymes. The method only relies on the first dimensional structures of homologous enzymes that can be obtained from genetic databases.  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