首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The primary structure of Escherichia coli L-threonine dehydrogenase   总被引:2,自引:0,他引:2  
The complete primary structures of Escherichia coli L-threonine dehydrogenase has been deduced by sequencing the cloned tdh gene. The primary structure so determined agrees with results obtained independently for the amino acid composition, the N-terminal amino acid sequence (20 residues), and a short sequence at the end of an internal peptide of the purified enzyme. The presence of a predicted Asp-Pro bond at residues 148 and 149 was confirmed by treatment of purified threonine dehydrogenase with dilute acid and subsequent analysis of the resulting cleavage products. The primary structure of L-threonine dehydrogenase from E. coli has been examined for possible homology to other NAD+-dependent dehydrogenases; indications are that this enzyme is a member of the zinc-containing long-chain alcohol/polyol dehydrogenase family.  相似文献   

2.
The gene encoding alanine dehydrogenase (AlaDH; EC 1.4.1.1) from the marine psychrophilic bacterium strain PA-43 was cloned, sequenced, and overexpressed in Escherichia coli. The primary structure was deduced on the basis of the nucleotide sequence. The enzyme subunit contains 371 amino acid residues, and the sequence is 90% and 77% identical, respectively, to AlaDHs from Shewanella Ac10 and Vibrio proteolyticus. The half-life of PA-43 AlaDH at 52 degrees C is 9 min, and it is thus more thermolabile than the AlaDH from Shewanella Ac10 or V. proteolyticus. The enzyme showed strong specificity for NAD(+) and l-alanine as substrates. The apparent K(m) for NAD(+) was temperature dependent (0.04 mM-0.23 mM from 15 degrees C to 55 degrees C). A comparison of the PA-43 deduced amino acid sequence to the solved three-dimensional structure of Phormidium lapideum AlaDH showed that there were likely to be fewer salt bridges in the PA-43 enzyme, which would increase enzyme flexibility and decrease thermostability. The hydrophobic surface character of the PA-43 enzyme was greater than that of P. lapideum AlaDH, by six residues. However, no particular modification or suite of modifications emerged as being clearly responsible for the psychrophilic character of PA-43 AlaDH.  相似文献   

3.
L-Threonine dehydrogenase, which forms aminoacetone from L-threonine and NAD, has been extensively purified from goat liver. A feedback inhibition of this enzyme has been observed with methylglyoxal. Kinetic data and other experiments indicate that methylglyoxal acts at a site other than the active site of the enzyme. The enzyme contains a single subunit of Mr 89,000. The apparent Km values of the enzyme for L-threonine and NAD were found to be 5.5 and 1 mM, respectively.  相似文献   

4.
Incubation of homogeneous preparations of L-threonine dehydrogenase from Escherichia coli with 2,3-butanedione, 2,3-pentanedione, phenylglyoxal, or 1,2-cyclohexanedione causes a time- and concentration-dependent loss of enzymatic activity; plots of log percent activity remaining versus time are linear to greater than 90% inactivation, indicative of pseudo-first order inactivation kinetics. The reaction order with respect to the concentration of modifying reagent is approximately 1.0 in each case suggesting that the loss of catalytic activity is due to one molecule of modifier reacting with each active unit of enzyme. Controls establish that this inactivation is not due to modifier-induced dissociation or photoinduced nonspecific alteration of the dehydrogenase. Essentially the same Km but decreased Vmax values are obtained when partially inactivated enzyme is compared with native. NADH (25 mM) and NAD+ (70 mM) give full protection against inactivation whereas much higher concentrations (i.e. 150 mM) of L-threonine or L-threonine amide provide a maximum of 80-85% protection. Amino acid analyses coupled with quantitative sulfhydryl group determinations show that enzyme inactivated 95% by 2,3-butanedione loses 7.5 arginine residues (out of 16 total)/enzyme subunit with no significant change in other amino acid residues. In contrast, only 2.4 arginine residues/subunit are modified in the presence of 80 mM NAD+. Analysis of the course of modification and inactivation by the statistical method of Tsou (Tsou, C.-L. (1962) Sci. Sin. 11, 1535-1558) demonstrates that inactivation of threonine dehydrogenase correlates with the loss of 1 "essential" arginine residue/subunit which quite likely is located in the NAD+/NADH binding site.  相似文献   

5.
When oxidized to cysteic acid by performic acid or converted to carboxymethylcysteine by alkylation of the reduced enzyme with iodoacetate, a total of six half-cystine residues/subunit are found in L-threonine dehydrogenase (L-threonine: NAD+ oxidoreductase, EC 1.1.1.103; L-threonine + NAD(+)----2-amino-3-oxobutyrate + NADH) from Escherichia coli K-12. Of this total, two exist in disulfide linkage, whereas four are titratable under denaturing conditions by dithiodipyridine, 5,5'-dithiobis(2-nitrobenzoic acid), or p-mercuribenzoate. The kinetics of enzyme inactivation and of modification by the latter two reagents indicate that threonine dehydrogenase has no free thiols that selectively react with bulky compounds. While incubation of the enzyme with a large excess of iodoacetamide causes less than 10% loss of activity, the native dehydrogenase is uniquely reactive with and completely inactivated by iodoacetate. The rate of carboxymethylation by iodoacetate of one -SH group/subunit is identical with the rate of inactivation and the carboxymethylated enzyme is no longer able to bind Mn2+. NADH (0.5 mM) provides 40% protection against this inactivation; 60 to 70% protection is seen in the presence of saturating levels of NADH plus L-threonine. Such results coupled with an analysis of the kinetics of inactivation caused by iodoacetate are interpreted as indicating the inhibitor first forms a reversible complex with a positively charged moiety in or near the microenvironment of a reactive -SH group in the enzyme before irreversible alkylation occurs. Specific alkylation of one -SH group/enzyme subunit apparently causes protein conformational changes that entail a loss of catalytic activity and the ability to bind Mn2+.  相似文献   

6.
7.
Gatzeva-Topalova PZ  May AP  Sousa MC 《Biochemistry》2004,43(42):13370-13379
Gram-negative bacteria including Escherichia coli, Salmonella typhimurium, and Pseudomonas aeruginosa can modify the structure of lipid A in their outer membrane with 4-amino-4-deoxy-l-arabinose (Ara4N). Such modification results in resistance to cationic antimicrobial peptides of the innate immune system and antibiotics such as polymyxin. ArnA is a key enzyme in the lipid A modification pathway, and its deletion abolishes both the Ara4N-lipid A modification and polymyxin resistance. ArnA is a bifunctional enzyme. It can catalyze (i) the NAD(+)-dependent decarboxylation of UDP-glucuronic acid to UDP-4-keto-arabinose and (ii) the N-10-formyltetrahydrofolate-dependent formylation of UDP-4-amino-4-deoxy-l-arabinose. We show that the NAD(+)-dependent decarboxylating activity is contained in the 360 amino acid C-terminal domain of ArnA. This domain is separable from the N-terminal fragment, and its activity is identical to that of the full-length enzyme. The crystal structure of the ArnA decarboxylase domain from E. coli is presented here. The structure confirms that the enzyme belongs to the short-chain dehydrogenase/reductase (SDR) family. On the basis of sequence and structure comparisons of the ArnA decarboxylase domain with other members of the short-chain dehydrogenase/reductase (SDR) family, we propose a binding model for NAD(+) and UDP-glucuronic acid and the involvement of residues T(432), Y(463), K(467), R(619), and S(433) in the mechanism of NAD(+)-dependent oxidation of the 4'-OH of the UDP-glucuronic acid and decarboxylation of the UDP-4-keto-glucuronic acid intermediate.  相似文献   

8.
The primary structure of bovine liver UDP-glucose dehydrogenase (UDPGDH), a hexameric, NAD(+)-linked enzyme, has been determined at the protein level. The 52-kDa subunits are composed of 468 amino acid residues, with a free N-terminus and a Ser/Asn microhetergeneity at one position. The sequence shares 29.6% positional identity with GDP-mannose dehydrogenase from Pseudomonas, confirming a similarity earlier noted between active site peptides. This degree of similarity is comparable to the 31.1% identity vs. the UDPGDH from type A Streptococcus. Database searching also revealed similarities to a hypothetical sequence from Salmonella typhimurium and to "UDP-N-acetyl-mannosaminuronic acid dehydrogenase" from Escherichia coli. Pairwise identities between bovine UDPGDH and each of these sequences were all in the range of approximately 26-34%. Multiple alignment of all 5 sequences indicates common ancestry for these 4-electron-transferring enzymes. There are 27 strictly conserved residues, including a cysteine residue at position 275, earlier identified by chemical modification as the expected catalytic residue of the second half-reaction (conversion of UDP-aldehydoglucose to UDP-glucuronic acid), and 2 lysine residues, at positions 219 and 338, one of which may be the expected catalytic residue for the first half-reaction (conversion of UDP-glucose to UDP-aldehydoglucose). A GXGXXG pattern characteristic of the coenzyme-binding fold is found at positions 11-16, close to the N-terminus as with "short-chain" alcohol dehydrogenases.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
We purified the psychrophilic and thermolabile malate dehydrogenase to homogeneity from a novel psychrotolerant, Flavobacterium frigidimaris KUC-1, isolated from Antarctic seawater. The enzyme was a homotetramer with a molecular weight of about 123 k and that of the subunit was about 32 k. The enzyme required NAD(P)(+) as a coenzyme and catalyzed the oxidation of L-malate and the reduction of oxalacetate specifically. The reaction proceeded through an ordered bi-bi mechanism. The enzyme was highly susceptible to heat treatment, and the half-life time at 40 degrees C was estimated to be 3.0 min. The k(cat)/K(m) (microM(-1).s(-1)) values for L-malate and NAD(+) at 30 degrees C were 289 and 2,790, respectively. The enzyme showed pro-R stereospecificity for hydrogen transfer at the C4 position of the nicotinamide moiety of the coenzyme. The enzyme contained 311 amino acid residues and much lower numbers of proline and arginine residues than other malate dehydrogenases.  相似文献   

10.
In Escherichia coli, Bacillus subtilis and Corynebacterium glutamicum, homoserine dehydrogenase (HD), the enzyme after the branch point of the threonine/methionine and lysine biosynthetic pathways, is allosterically inhibited by L-threonine. To investigate the regulation of the C. glutamicum HD enzyme by L-threonine, the structural gene, hom, was mutated by UV irradiation of whole cells to obtain a deregulated allele, homdr. L-Threonine inhibits the wild-type (wt) enzyme with a Ki of 0.16 mM. The deregulated enzyme remains 80% active in the presence of 50 mM L-threonine. The homdr gene mutant was isolated and cloned in E. coli. In a C. glutamicum wt host background, but not in E. coli, the cloned homdr gene is genetically unstable. The cloned homdr gene is overexpressed tenfold in C. glutamicum and is active in the presence of over 60 mM L-threonine. Sequence analysis revealed that the homdr mutation is a single nucleotide (G1964) deletion in codon 429 within the hom reading frame. The resulting frame-shift mutation radically alters the structure of the C terminus, resulting in ten amino acid (aa) changes and a deletion of the last 7 aa relative to the wt protein. These observations suggest that the C terminus may be associated with the L-threonine allosteric response. The homdr mutation is unstable and probably deleterious to the cell. This may explain why only one mutation was obtained despite repeated mutagenesis.  相似文献   

11.
12.
Synthesis of the biodegradative L-threonine dehydratase in Escherichia coli, Crookes strain, was prevented by dissolved oxygen concentrations of 6 micrometer or greater. This effect was shown to be exerted solely on synthesis, rather than being the result of enzyme inactivation in vivo. In addition to an anaerobic environment, maximum enzyme synthesis was dependent upon the presence of a complete complement of amino acids, with omission of L-threonine, L-valine, or L-leucine producing the largest decreases in enzyme formation. L-Threonine, the most essential of the amino acid requirements, could be partially replaced by DL-allothreonine or alpha-ketobutyrate. Half-maximal stimulation of enzyme synthesis occurred with 0.4 mM threonine in the medium. The roles of anaerobiosis and amino acids are interpreted as being in accord with the concept that threonine dehydratase functions in anaerobic energy production under conditions of amino acid sufficiency.  相似文献   

13.
A comparison of the primary structures of NAD(+)-dependent D-lactate dehydrogenase with L-lactate dehydrogenase and L-malate dehydrogenase failed to show any sequence similarity. However, D-2-hydroxyisocaproate dehydrogenase from Lactobacillus casei, glycerate dehydrogenase from cucumber, D-3-phosphoglycerate dehydrogenase and erythronate 4-phosphate dehydrogenase from Escherichia coli showed 38%, 24%, 24% and 22% amino acid identity, respectively. The profile analysis of the aligned sequences confirmed their relatedness. The hydropathy profiles of the aligned dehydrogenases were almost identical between residues 100-300 indicating largely preserved folding patterns of their polypeptide chains. The data suggest that L- and D-specific 2-hydroxy acid dehydrogenase genes evolved from two different ancestors and thus represent two different sets of enzyme families.  相似文献   

14.
L-Threonine dehydrogenase (TDH) from Escherichia coli is rapidly inactivated and develops a new absorbance peak at 347 nm when incubated with N-ethyl-5-phenylisoxazolium-3'-sulfonate (Woodward's reagent K, WRK). The cofactors, NAD+ or NADH (1.5 mM), provide complete protection against inactivation; L-threonine (60 mM) is approximately 50% as effective. Tryptic digestion of WRK-modified TDH followed by HPLC fractionation (pH 6.2) yields four 340-nm-absorbing peptides, two of which are absent from enzyme incubated with WRK and NAD+. Peptide I has the sequence TAICGTDVH (TDH residues 35-43), whereas peptide II is TAICGTDVHIY (residues 35-45). Peptides not protected are TMLDTMNHGGR (III, residues 248-258) and NCRGGRTHLCR (IV, residues 98-108). Absorbance spectra of these WRK-peptides were compared with WRK adducts of imidazole, 2-hydroxyethanethiolate, and acetate. Peptides III and IV have pH-dependent lambda max values (340-350 nm), consistent with histidine modification. Peptide I has pH-independent lambda max (350 nm) indicating that a thiol is modified. WRK, therefore, does not react specifically with carboxyl groups in this enzyme, but rather modifies Cys-38 in the active site of TDH; modification of His-105 and His-255 does not affect enzyme activity. These results are the first definitive proof of WRK modifying cysteine and histidine residues of a protein and show that enzyme inactivation by WRK associated with the appearance of new absorptivity at 340-350 nm does not establish modification of aspartate or glutamate residues, as has been assumed in numerous earlier reports.  相似文献   

15.
The crystal structure of a L-threonine dehydrogenase (L-ThrDH; EC 1.1.1.103) from the psychrophilic bacterium Flavobacterium frigidimaris KUC-1, which shows no sequence similarity to conventional L-ThrDHs, was determined in the presence of NAD and a substrate analog, glycerol. The asymmetric unit consisted of two subunits related by a two-fold rotation axis. Each monomer consisted of a Rossmann-fold domain and a carboxyl-terminal catalytic domain. The overall fold of F. frigidimaris L-ThrDH showed significant similarity to that of UDP-galactose 4-epimerase (GalE); however, structural comparison of the enzyme with E. coli and human GalEs showed clear topological differences in three loops (loop 1, loop 2 and the NAD-binding loop) around the substrate and NAD binding sites. In F. frigidimaris L-ThrDH, loops 1 and 2 insert toward the active site cavity, creating a barrier preventing the binding of UDP-glucose. Alternatively, loop 1 contributes to a unique substrate binding pocket in the F. frigidimaris enzyme. The NAD binding loop, which tightly holds the adenine ribose moiety of NAD in the Escherichia coli and human GalEs, is absent in F. frigidimaris L-ThrDH. Consequently, the cofactor binds to F. frigidimaris L-ThrDH in a reversible manner, unlike its binding to GalE. The substrate binding model suggests that the reaction proceeds through abstraction of the β-hydroxyl hydrogen of L-threonine via either a proton shuttle mechanism driven by Tyr143 and facilitated by Ser118 or direct proton transfer driven by Tyr143. The present structure provides a clear bench mark for distinguishing GalE-like L-ThrDHs from GalEs.  相似文献   

16.
We found N-methyl-L-amino acid dehydrogenase activity in various bacterial strains, such as Pseudomonas putida and Bacillus alvei, and cloned the gene from P. putida ATCC12633 into Escherichia coli. The enzyme purified to homogeneity from recombinant E. coli catalyzed the NADPH-dependent formation of N-alkyl-L-amino acids from the corresponding alpha-oxo acids (e.g. pyruvate, phenylpyruvate, and hydroxypyruvate) and alkylamines (e.g. methylamine, ethylamine, and propylamine). Ammonia was inert as a substrate, and the enzyme was clearly distinct from conventional NAD(P)-dependent amino acid dehydrogenases, such as alanine dehydrogenase (EC 1.4.1.1). NADPH was more than 300 times more efficient than NADH as a hydrogen donor in the enzymatic reductive amination. Primary structure analysis revealed that the enzyme belongs to a new NAD(P)-dependent oxidoreductase superfamily, the members of which show no sequence homology to conventional NAD(P)-dependent amino acid dehydrogenases and opine dehydrogenases.  相似文献   

17.
We found the occurrence of valine dehydrogenase in the cell extract of a psychrophilic bacterium, Cytophaga sp. KUC-1, isolated from Antarctic seawater and purified the enzyme to homogeneity. The molecular mass of the enzyme was determined to be approximately 154 kDa by gel filtration and that of the subunit was 43 kDa by SDS/PAGE: the enzyme was a homotetramer. The enzyme required NAD+ as a coenzyme, and catalyzed the oxidative deamination of L-valine, L-isoleucine, L-leucine and the reductive amination of alpha-ketoisovalerate, alpha-ketovalerate, alpha-ketoisocaproate, and alpha-ketocaproate. The reaction proceeds through an iso-ordered bi-bi mechanism. The enzyme was highly susceptible to heat treatment and the half-life at 45 degrees C was estimated to be 2.4 min. The kcat/Km (micro(-1).s(-1)) values for L-valine and NAD+ at 20 degrees C were 27.48 and 421.6, respectively. The enzyme showed pro-S stereospecificity for hydrogen transfer at the C4 position of the nicotinamide moiety of coenzyme. The gene encoding valine dehydrogenase was cloned into Escherichia coli (Novablue), and the primary structure of the enzyme was deduced on the basis of the nucleotide sequence of the gene encoding the enzyme. The enzyme contains 370 amino-acid residues, and is highly homologous with S. coelicolor ValDH (identity, 46.7%) and S. fradiae ValDH (43.1%). Cytophaga sp. KUC-1 ValDH contains much lower numbers of proline and arginine residues than those of other ValDHs. The changes probably lead to an increase in conformational flexibility of the Cytophaga enzyme molecule to enhance the catalytic activity at low temperatures.  相似文献   

18.
Bacterial UDP-glucose dehydrogenase (UDPGlcDH) is essential for formation of the antiphagocytic capsule that protects many virulent bacteria such as Streptococcus pyogenes andStreptococcus pneumoniae type 3 from the host's immune system. We have determined the X-ray structures of both native and Cys260Ser UDPGlcDH from S. pyogenes (74% similarity to S. pneumoniae) in ternary complexes with UDP-xylose/NAD(+) and UDP-glucuronic acid/NAD(H), respectively. The 402 residue homodimeric UDPGlcDH is composed of an N-terminal NAD(+) dinucleotide binding domain and a C-terminal UDP-sugar binding domain connected by a long (48 A) central alpha-helix. The first 290 residues of UDPGlcDH share structural homology with 6-phosphogluconate dehydrogenase, including conservation of an active site lysine and asparagine that are implicated in the enzyme mechanism. Also proposed to participate in the catalytic mechanism are a threonine and a glutamate that hydrogen bond to a conserved active site water molecule suitably positioned for general acid/base catalysis.  相似文献   

19.
The amino acid sequence is reported for CNBr and tryptic peptide fragments of the NAD(+)-dependent glutamate dehydrogenase of Clostridium symbiosum. Together with the N-terminal sequence, these make up about 75% of the total sequence. The sequence shows extensive similarity with that of the NADP(+)-dependent glutamate dehydrogenase of Escherichia coli (52% identical residues out of the 332 compared) allowing confident placing of the peptide fragments within the overall sequence. This demonstrated sequence similarity with the E. coli enzyme, despite different coenzyme specificity, is much greater than the similarity (31% identities) between the GDH's of C. symbiosum and Peptostreptococcus asaccharolyticus, both NAD(+)-linked. The evolutionary implications are discussed. In the 'fingerprint' region of the nucleotide binding fold the sequence Gly X Gly X X Ala is found, rather than Gly X Gly X X Gly. The sequence found here has previously been associated with NADP+ specificity and its finding in a strictly NAD(+)-dependent enzyme requires closer examination of the function of this structural motif.  相似文献   

20.
The gene encoding the NAD(+)-dependent glutamate dehydrogenase (GDH) of Clostridium symbiosum was cloned using the polymerase chain reaction (PCR) because it could not be recovered by standard techniques. The nucleotide sequence of the gdh gene was determined and it was overexpressed from the controllable tac promoter in Escherichia coli so that active clostridial GDH represented 20% of total cell protein. The recombinant plasmid complemented the nutritional lesion of an E. coli glutamate auxotroph. There was a marked difference between the nucleotide compositions of the coding region (G + C = 52%) and the flanking sequences (G + C = 30% and 37%). The structural gene encoded a polypeptide of 450 amino acid residues and relative molecular mass (M(r) 49,295 which corresponds to a single subunit of the hexameric enzyme. The DNA-derived amino acid sequence was consistent with a partial sequence from tryptic and cyanogen bromide peptides of the clostridial enzyme. The N-terminal amino acid sequence matched that of the purified protein, indicating that the initiating methionine is removed post-translationally, as in the natural host. The amino acid sequence is similar to those of other bacterial GDHs although it has a Gly-Xaa-Gly-Xaa-Xaa-Ala motif in the NAD(+)-binding domain, which is more typical of the NADP(+)-dependent enzymes. The sequence data now permit a detailed interpretation of the X-ray crystallographic structure of the enzyme and the cloning and expression of the clostridial gene will facilitate site-directed 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