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.
Aquaspillium arcticum is a psychrophilic bacterium that was isolated from arctic sediment and grows optimally at 4 degrees C. We have cloned, purified, and characterized malate dehydrogenase from A. arcticum (Aa MDH). We also have determined the crystal structures of apo-Aa MDH, Aa MDH.NADH binary complex, and Aa MDH.NAD.oxaloacetate ternary complex at 1.9-, 2.1-, and 2.5-A resolutions, respectively. The Aa MDH sequence is most closely related to the sequence of a thermophilic MDH from Thermus flavus (Tf MDH), showing 61% sequence identity and over 90% sequence similarity. Stability studies show that Aa MDH has a half-life of 10 min at 55 degrees C, whereas Tf MDH is fully active at 90 degrees C for 1 h. Aa MDH shows 2-3-fold higher catalytic efficiency compared with a mesophilic or a thermophilic MDH at the temperature range 4-10 degrees C. Structural comparison of Aa MDH and Tf MDH suggests that the increased relative flexibility of active site residues, favorable surface charge distribution for substrate and cofactor, and the reduced intersubunit ion pair interactions may be the major factors for the efficient catalytic activity of Aa MDH at low temperatures.  相似文献   

2.
The three-dimensional structure of four malate dehydrogenases (MDH) from thermophilic and mesophilic phototropic bacteria have been determined by X-ray crystallography and the corresponding structures compared. In contrast to the dimeric quaternary structure of most MDHs, these MDHs are tetramers and are structurally related to tetrameric malate dehydrogenases from Archaea and to lactate dehydrogenases. The tetramers are dimers of dimers, where the structures of each subunit and the dimers are similar to the dimeric malate dehydrogenases. The difference in optimal growth temperature of the corresponding organisms is relatively small, ranging from 32 to 55 degrees C. Nevertheless, on the basis of the four crystal structures, a number of factors that are likely to contribute to the relative thermostability in the present series have been identified. It appears from the results obtained, that the difference in thermostability between MDH from the mesophilic Chlorobium vibrioforme on one hand and from the moderate thermophile Chlorobium tepidum on the other hand is mainly due to the presence of polar residues that form additional hydrogen bonds within each subunit. Furthermore, for the even more thermostable Chloroflexus aurantiacus MDH, the use of charged residues to form additional ionic interactions across the dimer-dimer interface is favored. This enzyme has a favorable intercalation of His-Trp as well as additional aromatic contacts at the monomer-monomer interface in each dimer. A structural alignment of tetrameric and dimeric prokaryotic MDHs reveal that structural elements that differ among dimeric and tetrameric MDHs are located in a few loop regions.  相似文献   

3.
The thermostability in vitro of dimeric and tetrameric malate dehydrogenases [S)-malate:NAD+ oxidoreductase, EC 1.1.1.37) from mesophilic and thermophilic bacteria shows a good correlation to the growth temperature of the source organism but no consistent relationship to enzyme subunit structure. The thermophile malate dehydrogenases are, in general, more resistant to the surfactants, sodium dodecyl sulphate (SDS) and hexadecyltrimethylammonium bromide, and to the denaturants, guanidinium chloride and urea, than their mesophilic counterparts, with the dimer in each thermal class being more resistant to the chemical perturbants than the tetramer. Sedimentation analysis suggests that denaturation of the malate dehydrogenases by acid-periodate or SDS produces discrete subunits, whereas denaturation by guanidinium chloride followed by carboxymethylation yields ill-defined protein species. SDS and acid-periodate were therefore preferred to generate denatured malate dehydrogenases for use as immunogens and antigens. The native malate dehydrogenases exhibit immunological cross-reactivity only when they are in the same oligomeric form and derived from closely related species, which may, however, be from different thermal classes. Taking immunological cross-reactivity as an indicator of structural similarity, this supports the idea that the thermophilic trait evolved independently within each phyletic line. With denatured malate dehydrogenases as immunogens and antigens, cross-reactivity is manifested between all the malate dehydrogenases examined. This suggests that appreciable primary structural homology exists between the malate dehydrogenases, whether dimeric or tetrameric, from thermophiles and mesophiles and from various taxa.  相似文献   

4.
A cDNA clone, named ppcMDH-1 and covering a part of the coding region for the porcine cytosolic malate dehydrogenase (cMDH) mRNA, was isolated from a porcine liver cDNA library. Subsequently, mouse cMDH cDNA clones were isolated from mouse liver and heart cDNA libraries, using the ppcMDH-1 cDNA as a probe. The longest clone, named pmcMDH-5, was sequenced and the primary structure of the mouse cMDH deduced from its cDNA sequence showed that the mouse cMDH consists of the 334-amino acid residues. When the amino acid sequence of the mouse cMDH was compared with that of the porcine cMDH, they shared a 93% homology. On the other hand, the amino acid sequences of mouse cMDH and mitochondrial MDH (mMDH) showed about 23% overall homology. Surprisingly, comparison of the amino acid sequences among the mammalian and bacterial MDHs revealed that the homology between the mouse cMDH and thermophilic bacterial MDH, as well as the homology between the mouse mMDH and Escherichia coli MDH, markedly exceeds the intraspecies sequence homology between mMDH and cMDH from mice.  相似文献   

5.
The comparison of the three-dimensional structures of thermophilic (Thermus thermophilus) and mesophilic (Escherichia coli) 3-isopropylmalate dehydrogenases (IPMDH, EC 1.1.1.85) suggested that the existence of extra ion pairs in the thermophilic enzyme found in the intersubunit region may be an important factor for thermostability. As a test of our assumption, glutamine 200 in the E. coli enzyme was turned into glutamate (Q200E mutant) to mimic the thermophilic enzyme at this site by creating an intersubunit ion pair which can join existing ion clusters. At the same site in the thermophilic enzyme we changed glutamate 190 into glutamine (E190Q), hereby removing the corresponding ion pair. These single amino acid replacements resulted in increased thermostability of the mesophilic and decreased thermostability of the thermophilic enzyme, as measured by spectropolarimetry and differential scanning microcalorimetry.  相似文献   

6.
A 3 kb DNA fragment containing the gene (mdh) encoding malate dehydrogenase (MDH) from the thermophile Thermus aquaticus B was cloned in Escherichia coli and its nucleotide sequence determined. Comparative analysis showed the nucleotide sequence to be very closely related to that determined for the Thermus flavus mdh gene and flanking regions, with no differences between the predicted amino acid sequences of the MDHs. A proximal open reading frame, identified as the sucD gene, and the mdh gene may be parts of the same operon in T. aquaticus B. Expression of the T. aquaticus B mdh gene in E. coli was found to be at a relatively low level. A simple method for purification of thermostable MDH from the E. coli clone containing the T. aquaticus B mdh gene is presented.  相似文献   

7.
Malate dehydrogenases from Streptosporangium roseum (DSM 43021), Planomonospora venezuelensis (DSM 43178), Microtetraspora glauca (ATCC 23057), Actinoplanes missouriensis (DSM 43046), Streptomyces atratus (ATCC 14046), Kibdelosporangium aridum (ATCC 39323), and from Phenylobacterium immobile, strain E (DSM 1986) were purified to homogeneity. The N-terminal amino-acid sequences were determined and compared with known prokaryotic and eukaryotic sequence data. The partial sequences from Actinomycetales enzymes include a string of amino acids which is also present in the N-terminal region of malate dehydrogenases from Thermus flavus and from mammalian cytoplasm.  相似文献   

8.
Genetic and cytogenetic locations of the structural genes for the NAD-dependent malate dehydrogenases have been studied. The mitochondrial form (mMDH) is coded for by a gene (mMdh) found at 62.6 on the third chromosome and included in Df(3R)R14, which includes 90C2-91A3 in the salivary gland chromosomes. Based on its inclusion within several J (Jammed; 2-41.0) deficiencies, the structural gene (cMdh) for the cytoplasmic form (cMDH) was determined to lie in region 31B-E, confirming the earlier finding of Grell. Flies lacking any cMDH activity (cMdhn-gamma 10069/Df(2L)J-der-27) were both viable and fertile.  相似文献   

9.
During P deficiency, the increased activity of malate dehydrogenase (MDH, EC 1.1.1.37) can lead to malate accumulation. Cytosolic- and nodule-enhanced MDH (cMDH and neMDH, respectively) are known isoforms, which contribute to MDH activity in root nodules. The aim of this study was to investigate the role of the cMDH isoforms in nodule malate supply under P deficiency. Nodulated lupins (Lupinus angustifolius var. Tanjil) were hydroponically grown at adequate P (+P) or low P (−P). Total P concentration in nodules decreased under P deficiency, which coincided with an increase in total MDH activity. A consequence of higher MDH activity was the enhanced accumulation of malate derived from dark CO2 fixation via PEPC and not from pyruvate. Although no measurable neMDH presence could be detected via PCR, gene-specific primers detected two 1 kb amplicons of cMDH, designated LangMDH1 (corresponding to +P, HQ690186) and LangMDH2 (corresponding to −P, HQ690187), respectively. Sequencing analyses of these cMDH amplicons showed them to be 96% identical on an amino acid level. There was a high degree of diversification between proteins detected in this study and other known MDH proteins, particularly those from other leguminous plants. Enhanced malate synthesis in P-deficient nodules was achieved via increased anaplerotic CO2 fixation and subsequent higher MDH activities. Novel isoforms of cytosolic MDH may be involved, as shown by gene expression of specific genes under P deficiency.  相似文献   

10.
The amino acid sequences of two thermophilic and five mesophilic glyceraldehyde-3-phosphate dehydrogenases have been compared with the known three-dimensional structure of this enzyme to determine the factors responsible for thermal stability. The changes are greatest in the S-loop regions at the center of the tetramer, which show a quantitative increase in hydrophobicity and polarity that can strengthen subunit interactions in a complementary manner. The S-loops also show increases in residue volume and bulk that may indicate a tighter packing at the molecular center. In addition, there are changes in the secondary structural parameters indicating that the helices, in particular, may be more stable in the thermophilic proteins. Increases in the hydrophobicity of domain and subunit contacts for the Thermus aquaticus glyceraldehyde-3-phosphate dehydrogenase may explain why it is the most thermostable protein in this series.  相似文献   

11.
The nucleotide sequence of the malate dehydrogenase (mdh) gene from a thermophilic bacterium, Thermus flavus, was determined. The amino acid sequence of the Thermus malate dehydrogenase resembled that of the porcine heart cytoplasmic enzyme to a certain extent, and Asp-159 and His-187 were identified as possible essential residues for the catalytic function. The mutated mdh gene was also cloned from a spontaneous mutant of T. flavus containing a higher activity of the enzyme. Its mutation point was determined to be a single nucleotide exchange from C to T which caused Thr-190 to be substituted by isoleucine. The mutated enzyme showed resistance to substrate inhibition, an increase in both kcat and Km, and a shift toward a more acid optimum pH for the enzyme reaction.  相似文献   

12.
Malate synthases from a thermophilic Bacillus and Escherichia coli have been isolated in a high state of purity. Molecular weights of these two proteins determined in the native state and after denaturation in sodium dodecyl sulfate-mercaptoethanol show that the enzymes are monomeric. This conclusion is supported, for the thermophile enzyme, by the result of an electrophoretic analysis of that protein after treatment with dimethylsuberimidate and denaturation. The thermophilic Bacillus malate synthase is considerably more thermostable than its mesophilic counterparts from E. coli, Bacillus licheniformis, and Pseudomonas indigofera. It is, however, markedly labilized by an increase in the ionic strength of the medium brought about by the addition of 0.2 M potassium chloride or in pH above 9. Increased ionic strength has little effect on the thermostability of the mesophilic bacterial malate synthases. These observations provide strong support for the idea that monomeric proteins in thermophiles owe their unusual heat stability to the presence of salt bridges in their tertiary structure.  相似文献   

13.
Tertiary RNA structures from thermophilic bacteria generally are more stable than their mesophilic homologs. To understand the structural basis of the increase in stability, we investigated equilibrium folding of the specificity domain (S-domain) of RNase P RNA from a mesophilic (Escherichia coli) and a thermophilic (Thermus thermophilus) bacterium. Equilibrium folding of both S-domains is described by a minimal, three-state folding scheme, U-to-I-to-N. In the I-to-N transition of the thermophilic S-domain, more structure forms and protections are stronger against T1 nuclease and hydroxyl radical reactions. Phylogenetic comparison in the context of the native structure reveals that among 39 nucleotide differences between these S-domains, 12 likely contribute to higher stability. These residues participate in extensive networks of hydrogen bonding, stacking, and metal ion coordination throughout the molecule. The thermophilic S-domain achieves higher stability by mutating strategic base pairs to G-C, decreasing surface accessibility of the native state, and increasing the amount of structure formation in the native folding transition. An E. coli S-domain mutant containing these 12 nt has the same stability and folding cooperativity as the T. thermophilus S-domain. E. coli S-domain mutants containing a subset of 4 or 6 nt have the same stability as the T. thermophilus S-domain but the same folding cooperativity as the E. coli S-domain. These results show that increasing stability can be accomplished by mutations within a local structure, but increasing folding cooperativity needs concerted changes among multiple structural units.  相似文献   

14.
Kirby RR 《Gene》2000,245(1):81-88
The evolutionary history of the malate dehydrogenase (MDH) gene family [NAD-dependent MDH; EC 1.1.1.37 and NAD(P)-dependent MDH; EC 1.1.1.82] has received much attention. MDHs have also featured extensively as electrophoretic markers in population genetics and evolutionary ecology, and in many cases, intraspecific variation in MDH has been correlated with environmental variables. However, while the amino acid residues essential for MDH function are known, no studies have examined intraspecific nucleotide variation despite evidence indicating that natural selection may be operating on this locus. This study presents two sets of degenerate oligonucleotide PCR primers to facilitate the cloning of cytosolic MDH (cMDH) and mitochondrial MDH (mMDH) from a broad range of animals (cMDH) and animals and plants (mMDH). These primers were used to obtain putative cMDH and mMDH cDNAs from the mollusc Nucella lapillus. The N. lapillus cMDH cDNA was found to encode a putative cMDH protein of 334aa and 36kDa, while the mMDH cDNA encoded a putative mature mMDH protein of 315aa and 33kDa. The putative amino acid sequences of the two compartmentalised N. lapillus MDHs are presented and compared to other known MDH sequences.  相似文献   

15.
Malate dehydrogenase (MDH) catalyzes the readily reversible reaction of oxaloacetate ; malate using either NADH or NADPH as a reductant. In plants, the enzyme is important in providing malate for C 4 metabolism, pH balance, stomatal and pulvinal movement, respiration, β-oxidation of fatty acids, and legume root nodule functioning. Due to its diverse roles the enzyme occurs as numerous isozymes in various organelles. While antibodies have been produced and cDNAs characterized for plant mitochondrial, glyoxysomal, and chloroplast forms of MDH, little is known of other forms. Here we report the cloning and characterization of cDNAs encoding five different forms of alfalfa MDH, including a plant cytosolic MDH (cMDH) and a unique novel nodule-enhanced MDH (neMDH). Phylogenetic analyses show that neMDH is related to mitochondrial and glyoxysomal MDHs, but diverge from these forms early in land plant evolution. Four of the five forms could effectively complement an E. coli Mdh mutant. RNA and protein blots show that neMDH is most highly expressed in effective root nodules. Immunoprecipitation experiments show that antibodies produced to cMDH and neMDH are immunologically distinct and that the neMDH form comprises the major form of total MDH activity and protein in root nodules. Kinetic analysis showed that neMDH has a turnover rate and specificity constant that can account for the extraordinarily high synthesis of malate in nodules.   相似文献   

16.
The complete gene of cytosolic malate dehydrogenase (cMDH) from Camellia sinensis, called Cs cMDH, was obtained by RT PCR and rapid amplification of cDNA ends (GenBank accession number GQ845406). This gene was 1 235 bp in length, encoding a protein of 332 amino acids with the putative molecular weight of 355 kD. The Ecoli Rosetta (DE3) harboring pGEX MDH was induced by 05 mmol·L 1 IPTG at 32℃ for 3 hours, and a 615 kD glutathione Stransferase (GST) fused MDH was obtained in soluble form. The results of NCBI BLAST revealed that Cs cMDH shared 88%-93% of amino acid sequence identity with other cMDH from different higher plants. According to the multiple sequence alignment based on the three dimensional structure of protein, Cs cMDH was predicted to be a dimer with thirteen β sheet and thirteen α helix of each subunit. Cs cMDH contains typical fingerprint sequence (G12AAGQIG18) as all MDHs. The amino acid D43 in Cs cMDH is conserved in all NAD MDHs. Cs cMDH also has some conserved sequence units homologous to other NAD MDHs, such as NAD+ binding sites, catalytic motif and substrate binding sites. Moreover, Cs cMDH contains six Cys which are highly conserved in all plant NAD cMDHs. Therefore, Cs cMDH was inferred to be NAD dependent cMDH. The present study may provide the fundament for the further functional characterization of Cs cMDH.  相似文献   

17.
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   

18.
The heat treatment of recombinant mesophilic cells having heterologous thermophilic enzymes results in the denaturation of indigenous mesophilic enzymes and the elimination of undesired side reactions; therefore, highly selective whole-cell catalysts comparable to purified enzymes can be readily prepared. However, the thermolysis of host cells leads to the heat-induced leakage of thermophilic enzymes, which are produced as soluble proteins, limiting the exploitation of their excellent stability in repeated and continuous reactions. In this study, Escherichia coli cells having the thermophilic fumarase from Thermus thermophilus (TtFTA) were treated with glutaraldehyde to prevent the heat-induced leakage of the enzyme, and the resulting cells were used as a whole-cell catalyst in repeated and continuous reactions. Interestingly, although electron microscopic observations revealed that the cellular structure of glutaraldehyde-treated E. coli was not apparently changed by the heat treatment, the membrane permeability of the heated cells to relatively small molecules (up to at least 3 kDa) was significantly improved. By applying the glutaraldehyde-treated E. coli having TtFTA to a continuous reactor equipped with a cell-separation membrane filter, the enzymatic hydration of fumarate to malate could be operated for more than 600 min with a molar conversion yield of 60% or higher.  相似文献   

19.
The steady-state kinetics of alcohol dehydrogenases (alcohol:NAD+ oxidoreductase, EC 1.1.1.1 and alcohol:NADP+ oxidoreductase, EC 1.1.1.2), lactate dehydrogenases (l-lactate:NAD+ oxidoreductase, EC 1.1.1.27 and d-lactate:NAD+ oxidoreductase, EC 1.1.1.28), malate dehydrogenase (l-malate:NAD+ oxidoreductase, EC 1.1.1.37), and glyceraldehyde-3-phosphate dehydrogenases [d-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12] from different sources (prokaryote and eukaryote, mesophilic and thermophilic organisms) have been studied using NAD(H), N6-(2-carboxyethyl)-NAD(H), and poly(ethylene glycol)-bound NAD(H) as coenzymes. The kinetic constants for NAD(H) were changed by carboxyethylation of the 6-amino group of the adenine ring and by conversion to macromolecular form. Enzymes from thermophilic bacteria showed especially high activities for the derivatives. The relative values of the maximum velocity (NAD = 1) of Thermus thermophilus malate dehydrogenase for N6-(2-carboxyethyl)-NAD and poly(ethylene glycol)-bound NAD were 5.7 and 1.9, respectively, and that of Bacillus stearothermophilus glyceraldehyde-3-phosphate dehydrogenase for poly(ethylene glycol)-bound NAD was 1.9.  相似文献   

20.
Thermostable dimeric malate dehydrogenase (MDH) was isolated from the microorganism of hydrothermal vents Vulcanithermus medioatlanticus. The enzyme was electrophoretically homogeneous and possessed the specific activity of 6.9 U/mg. The large molecular weight of the subunits (55 kD) is likely to provide the rigidity of the enzyme structure (the activation energy of the enzymatic reaction is 32.6 kJ/mol). The thermophilic MDH differs little from the mesophilic enzyme in terms of kinetic and regulatory characteristics.  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