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.
In our previous study, we showed that a chimeric isopropylmalate dehydrogenase, 2T2M6T, between an extreme thermophile, Thermus thermophilus, and a mesophile, Bacillus subtilis, isopropylmalate dehydrogenases (the name roughly denotes the primary structure; the first 20% from the N-terminal is coded by the thermophile leuB gene, next 20% by mesophile, and the rest by the thermophile gene) denatured in two steps with a stable intermediate, suggesting that in the chimera some of the interdomain interaction was lost by amino acid substitutions in the "2M" part. To identify the residues involved in the interdomain interactions, the first and the second halves of the 2M part of the chimera were substituted with the corresponding sequence of the thermophile enzyme. Both chimeras, 3T1M6T and 2T1M7T, apparently showed one transition in the thermal denaturation without any stable intermediate state, suggesting that the cooperativity of the conformational stability was at least partly restored by the substitutions. The present study also suggested involvement of one or more basic residues in the unusual stability of the thermophile enzyme. Received: September 29, 1998 / Accepted: June 25, 1999  相似文献   

2.
The gene coding for isocitrate dehydrogenase of an extreme thermophile, Thermus thermophilus HB8, was cloned and sequenced. This gene consists of a single open reading frame of 1,485 bp preceded by a Shine-Dalgarno ribosome binding site. Promoter- and terminatorlike sequences were detected upstream and downstream of the open reading frame, respectively. The G + C content of the coding region was 65.6%, and that of the third nucleotide of the codons was 90.3%. On the basis of the deduced amino acid sequence, the Mr of the monomeric enzyme was calculated as 54,189, an Mr which is similar to that of the purified protein determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. A comparison of the amino acid sequence of the T. thermophilus enzyme with that of the Escherichia coli enzyme showed (i) a 37% overall similarity; (ii) the conservation of the Ser residue, which is known to be phosphorylated in the E. coli enzyme, and of the surrounding sequence; and (iii) the presence of 141 extra residues at the C terminus of the T. thermophilus enzyme. T. thermophilus isocitrate dehydrogenase showed a high sequence homology (33% of the amino acid sequence is identical) to isopropylmalate dehydrogenase from the same organism and was suggested to have evolved from a common ancestral enzyme.  相似文献   

3.
The gene coding for isocitrate dehydrogenase of an extreme thermophile, Thermus thermophilus HB8, was cloned and sequenced. This gene consists of a single open reading frame of 1,485 bp preceded by a Shine-Dalgarno ribosome binding site. Promoter- and terminatorlike sequences were detected upstream and downstream of the open reading frame, respectively. The G + C content of the coding region was 65.6%, and that of the third nucleotide of the codons was 90.3%. On the basis of the deduced amino acid sequence, the Mr of the monomeric enzyme was calculated as 54,189, an Mr which is similar to that of the purified protein determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. A comparison of the amino acid sequence of the T. thermophilus enzyme with that of the Escherichia coli enzyme showed (i) a 37% overall similarity; (ii) the conservation of the Ser residue, which is known to be phosphorylated in the E. coli enzyme, and of the surrounding sequence; and (iii) the presence of 141 extra residues at the C terminus of the T. thermophilus enzyme. T. thermophilus isocitrate dehydrogenase showed a high sequence homology (33% of the amino acid sequence is identical) to isopropylmalate dehydrogenase from the same organism and was suggested to have evolved from a common ancestral enzyme.  相似文献   

4.
Thermal stability of the Thermus thermophilus isopropylmalate dehydrogenase enzyme was substantially lost upon the deletion of three residues from the C-terminus. However, the stability was partly recovered by the addition of two, four and seven amino acid residues (called HD177, HD708 and HD711, respectively) to the C-terminal region of the truncated enzyme. Three structures of these mutant enzymes were determined by an X-ray diffraction method. All protein crystals belong to space group P2(1) and their structures were solved by a standard molecular replacement method where the original dimer structure of the A172L mutant was used as a search model. Thermal stability of these mutant enzymes is discussed based on the 3D structure with special attention to the width of the active-site groove and the minor groove, distortion of beta-sheet pillar structure and size of cavity in the domain-domain interface around the C-terminus. Our previous studies revealed that the thermal stability of isopropylmalate dehydrogenase increases when the active-site cleft is closed (the closed form). In the present study it is shown that the active-site cleft can be regulated by open-close movement of the minor groove located at the opposite side to the active-site groove on the same subunit, through a paperclip-like motion.  相似文献   

5.
A temperature-jump (T-jump) time-resolved X-ray crystallographic technique using the Laue method was developed to detect small, localized structural changes of proteins in crystals exposed to a temperature increase induced by laser irradiation. In a chimeric protein between thermophilic and mesophilic 3-isopropylmalate dehydrogenases (2T2M6T), the initial structural change upon T-jump to a denaturing temperature (approximately 90 degrees C) was found to be localized at a region which includes a beta-turn and a loop located between the two domains of the enzyme. A mutant, 2T2M6T-E110P/S111G/S113E, having amino acid replacements in this beta-turn region with the corresponding residues of the thermophilic enzyme, showed greater stability than the original chimera (increase of T:(m) by approximately 10 degrees C) and no T-jump-induced structural change in this region was detected by our method. These results indicate that thermal unfolding of the original chimeric enzyme, 2T2M6T, is triggered in this beta-turn region.  相似文献   

6.
A new type of chimeric 3-isopropylmalate dehydrogenase (2T2M6T) was produced by expressing the fused gene of Bacillus subtilis and Thermus thermophilus. The enzyme shows heat stability intermediate between those of the parents. The crystal of the enzyme belongs to the space group of P3(2)21, with cell dimensions of a = b = 78.9 A and c = 158.9 A. Two thermostable mutants of the chimeric enzyme were prepared by site-directed mutagenesis and then crystallized.  相似文献   

7.
Random mutagenesis of Thermus thermophilus 3-isopropylmalate dehydrogenase revealed that a substitution of Val126Met in a hinge region caused a marked increase in specific activity, particularly at low temperatures, although the site is far from the binding residues for 3-isopropylmalate and NAD. To understand the molecular mechanism, residue 126 was substituted with one of eight other residues, Gly, Ala, Ser, Thr, Glu, Leu, Ile or Phe. Circular dichroism analyses revealed a decreased thermal stability of the mutants (Delta T ((1/2))= 0-13 degrees C), indicating structural perturbations caused by steric conflict with surrounding residues having larger side chains. Kinetic parameters, k(cat) and K(m) values for isopropylmalate and NAD, were also affected by the mutation, but the resulting k(cat)/K(m) values were similar to that of the wild-type enzyme, suggesting that the change in the catalytic property is caused by the change in free-energy level of the Michaelis complex state relative to that of the initial state. The kinetic parameters and activation enthalpy change (Delta H (double dagger)) showed good correlation with the van der Waals volume of residue 126. These results suggested that the artificial cold adaptation (enhancement of k(cat) value at low temperatures) resulted from the destabilization of the ternary complex caused by the increase in the volume of the residue at position 126.  相似文献   

8.
The sequence and resolved structure of thermotrophic isopropylmalate dehydrogenase (IPMDH) and a related protein, mesotrophic isocitrate dehydrogenase (IDH), were compared emphasizing clusters of charged residues identified from X-ray crystallographic studies (Wallon, G., Kryger, G., Lovett, S. T., Oshima, T., Ringe, D., and Petsko, G. A. (1997) J. Mol. Biol. 266, 1016-1031). Mesotrophic isocitrate dehydrogenase was used for comparison because crystallographic data for a mesotrophic form of IPMDH was not available in the database. The structural features in the region of these clusters were compared and localized conformational differences were found in the thermotroph compared to the mesotroph. Because the overall topology of the two proteins is similar, it was concluded that these localized structural differences induced by electrostatic interactions between charged residues in the thermotrophic enzyme were responsible for the enhanced thermal stability of proteins from thermotroph.  相似文献   

9.
Doyle SA  Beernink PT  Koshland DE 《Biochemistry》2001,40(14):4234-4241
Isocitrate dehydrogenase (IDH) catalyzes the oxidative decarboxylation of isocitrate and has negligible activity toward other (R)-malate-type substrates. The S113E mutant of IDH significantly improves its ability to utilize isopropylmalate as a substrate and switches the substrate specificity (k(cat)/K(M)) from isocitrate to isopropylmalate. To understand the structural basis for this switch in substrate specificity, we have determined the crystal structure of IDH S113E in a complex with isopropylmalate, NADP, and Mg(2+) to 2.0 A resolution. On the basis of a comparison with previously determined structures, we identify distinct changes caused by the amino acid substitution and by the binding of substrates. The S113E complex exhibits alterations in global and active site conformations compared with other IDH structures that include loop and helix conformational changes near the active site. In addition, the angle of the hinge that relates the two domains was altered in this structure, which suggests that the S113E substitution and the binding of substrates act together to promote catalysis of isopropylmalate. Ligand binding results in reorientation of the active site helix that contains residues 113 through 116. E113 exhibits new interactions, including van der Waals contacts with the isopropyl group of isopropylmalate and a hydrogen bond with N115, which in turn forms a hydrogen bond with NADP. In addition, the loop and helix regions that bind NADP are altered, as is the loop that connects the NADP binding region to the active site helix, changing the relationship between substrates and enzyme. In combination, these interactions appear to provide the basis for the switch in substrate specificity.  相似文献   

10.
1. D-Glyceraldehyde-3-phosphate dehydrogenase from an extreme thermophile, T. thermophilus strain HB8, was purified and crystallized. 2. The enzyme was found to possess remarkable heat stability, being slowly inactivated at 90 degrees C. 3. Basic kinetic constants and pH profile are reported. The enzyme was activated 25-fold by 90 mM NH4Cl, and also by ethanol up to 5-fold at 30 degrees C. 4. The enzyme was found to be far more resistant to urea or sodium dodecylsulfate than the rabbit enzyme. 5. The enzyme was shown to be a tetramer of molecular weight 130000--135000. Amino acid composition analysis revealed no unusual features. Circular dichroic spectra suggested that the contents of the ordered structure of the thermophile enzyme are similar to those of the rabbit enzyme. 6. The other catalytic properties of the thermophile enzyme are discussed in comparison with those of the enzymes from other sources.  相似文献   

11.
The thermal unfolding process of a chimeric 3-isopropylmalate dehydrogenase made of parts from an extreme thermophile, Thermus thermophilus, and a mesophile, Bacillus subtilis, enzymes was studied by CD spectrophotometry and differential scanning calorimetry (DSC). The enzyme is a homodimer with a subunit containing two structural domains. The DSC melting profile of the chimeric enzyme in 20 mM NaHCO3, pH 10.4, showed two endothermic peaks, whereas that of the T. thermophilus wild-type enzyme had one peak. The CD melting profiles of the chimeric enzyme under the same conditions as the DSC measurement, also indicated biphasic unfolding transition. Concentration dependence of the unfolding profile revealed that the first phase was protein concentration-independent, whereas the second transition was protein concentration-dependent. When cooled after the first transition, the intermediate was isolated, which showed only the second transition upon heating. These results indicated the existence of a stable dimeric intermediate followed by the further unfolding and dissociation in the thermal unfolding of the chimeric enzyme at pH 10-11. Because the portion derived from the mesophilic isopropylmalate dehydrogenase in the chimeric enzyme is located in the hinge region between two domains of the enzyme, it is probably responsible for weakening of the interdomain interaction and causing the decooperativity of two domains. The dimeric form of the intermediate suggested that the first unfolding transition corresponds to the unfolding of domain 1 containing the N- and C-termini of the enzyme, and the second to that of domain 2 containing the subunit interface.  相似文献   

12.
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   

13.
The polyol dehydrogenase PDH-11300 from Deinococcus geothermalis was cloned, functionally expressed in Escherichia coli and biochemically characterized. The enzyme showed the highest activity in the oxidation of xylitol and 1,2-hexanediol and had an optimum temperature of 45°C. The enzyme exhibited a T(50)(60)-value of 48.3°C. The T(50)(60) is the temperature where 50% of the initial activity remains after incubation for 1h. In order to elucidate the structural reasons contributing to thermostability, the substrate-binding loop of PDH-11300 was substituted by the loop-region of a homolog enzyme, the galactitol dehydrogenase from Rhodobacter sphaeroides (PDH-158), resulting in a chimeric enzyme (PDH-loop). The substrate scope of this chimera basically represented the average of both wild-type enzymes, but surprisingly the T(50)(60) was noticeably increased by 7°C up to 55.3°C. Further mutations in the active site led to identification of residues crucial for enzyme activity. The cofactor specificity was successfully altered from NADH to NADPH by an Asp55Asn mutation, which is located at the NAD(+) binding cleft, without influencing the catalytic properties of the dehydrogenase.  相似文献   

14.
Although enzymes of thermophilic organisms are often very resistant to thermal denaturation, they are usually less active than their mesophilic or psychrophilic homologues at moderate or low temperatures. To explore the structural features that would improve the activity of a thermophilic enzyme at less than optimal temperatures, we randomly mutated the DNA of single-site mutants of the thermostable Thermus thermophilus 3-isopropylmalate dehydrogenase that already had improved low-temperature activity and selected for additional improved low-temperature activity. A mutant (Ile279 → Val) with improved low-temperature activity contained a residue that directly interacts with the adenine of the coenzyme NAD(+), suggesting that modulation of the coenzyme-binding pocket's volume can enhance low-temperature activity. This idea was further supported by a saturation mutagenesis study of the two codons of two other residues that interact with the adenine. Furthermore, a similar type of amino acid substitution also improved the catalytic efficiency of another thermophilic dehydrogenase, T. thermophilus lactate dehydrogenase. Steady-state kinetic experiments showed that the mutations all favorably affected the catalytic turnover numbers. Thermal stability measurements demonstrated that the mutants remain very resistant to heat. Calculation of the energetic contributions to catalysis indicated that the increased turnover numbers are the result of destabilized enzyme-substrate-coenzyme complexes. Therefore, small changes in the side chain volumes of coenzyme-binding residues improved the catalytic efficiencies of two thermophilic dehydrogenases while preserving their high thermal stabilities and may be a way to improve low-temperature activities of dehydrogenases in general.  相似文献   

15.
We improved the thermal stability of 3-isopropylmalate dehydrogenase from Bacillus subtilis by an in vivo evolutionary technique using an extreme thermophile, Thermus thermophilus, as a host cell. The leuB gene encoding B. subtilis 3-isopropylmalate dehydrogenase was integrated into the chromosome of a leuB-deficient strain of T. thermophilus. The resulting transformant showed a leucine-autotrophy at 56 degrees C but not at 61 degrees C and above. Phenotypically thermostabilized strains that can grow at 61 degrees C without leucine were isolated from spontaneous mutants. Screening temperature was stepwise increased from 61 to 66 and then to 70 degrees C and mutants that showed a leucine-autotrophic growth at 70 degrees C were obtained. DNA sequence analyses of the leuB genes from the mutant strains revealed three stepwise amino acid replacements, threonine-308 to isoleucine, isoleucine-95 to leucine, and methionine-292 to isoleucine. The mutant enzymes with these amino acid replacements were more stable against heat treatment than the wild-type enzyme. Furthermore, the triple-mutant enzyme showed significantly higher specific activity than that of the wild-type enzyme.  相似文献   

16.
Threo-Ds-3-isopropylmalate dehydrogenase coded by the leuB gene from an extreme thermophile, Thermus thermophilus strain HB8, was expressed in Escherichia coli carrying a recombinant plasmid. The thermostable enzyme thus produced was extracted from the E. coli cells, purified, and crystallized. The enzyme was shown to be a dimer of identical subunits of molecular weight (4.0 +/- 0.5) x 10(4). The Km for threo-Ds-3-isopropylmalate was estimated to be 8.0 x 10(-5) M and that for NAD 6.3 x 10(-4) M. The optimum pH at 75 degrees C in the presence of 1.2 M KCl was around 7.2. The presence of Mg2+ or Mn2+ was essential for the enzyme action. The enzyme was activated about 30-fold by the addition of 1 M KCl or RbCl. The high salt concentration decelerated the thermal unfolding of the enzyme, and accelerated the aggregation of the unfolded protein. Based on these effects, the molecular mechanism of the unusual stability of the enzyme is discussed.  相似文献   

17.
We compared cleavage efficiencies of mono-molecular and bipartite model RNAs as substrates for RNase P RNAs (M1 RNAs) and holoenzymes from E. coli and Thermus thermophilus, an extreme thermophilic eubacterium. Acceptor stem and T arm of pre-tRNA substrates are essential recognition elements for both enzymes. Impairing coaxial stacking of acceptor and T stems and omitting the T loop led to reduced cleavage efficiencies. Small model substrates were less efficiently cleaved by M1 RNA and RNase P from T. thermophilus than by the corresponding E. coli activities. Competition kinetics and gel retardation studies showed that truncated tRNA substrates are less tightly bound by RNase P and M1 RNA from both bacteria. Our data further indicate that (pre-)tRNA interacts stronger with E. coli than T. thermophilus M1 RNA. Thus, low cleavage efficiencies of truncated model substrates by T. thermophilus RNase P or M1 RNA could be explained by a critical loss of important contact points between enzyme and substrate. In addition, acceptor stem--T arm substrates, composed of two synthetic RNA fragments, have been designed to mimic internal cleavage of any target RNA molecule available for base pairing.  相似文献   

18.
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   

19.
The crystal structures of aspartyl-tRNA synthetase (AspRS) from Thermus thermophilus, a prokaryotic class IIb enzyme, complexed with tRNA(Asp) from either T. thermophilus or Escherichia coli reveal a potential intermediate of the recognition process. The tRNA is positioned on the enzyme such that it cannot be aminoacylated but adopts an overall conformation similar to that observed in active complexes. While the anticodon loop binds to the N-terminal domain of the enzyme in a manner similar to that of the related active complexes, its aminoacyl acceptor arm remains at the entrance of the active site, stabilized in its intermediate conformational state by non-specific interactions with the insertion and catalytic domains. The thermophilic nature of the enzyme, which manifests itself in a very low kinetic efficiency at 17 degrees C, the temperature at which the crystals were grown, is in agreement with the relative stability of this non-productive conformational state. Based on these data, a pathway for tRNA binding and recognition is proposed.  相似文献   

20.
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   

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