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 62 毫秒
1.
Diaminopimelate (DAP) epimerase is involved in the biosynthesis of meso-DAP and lysine, which are important precursors for the synthesis of peptidoglycan, housekeeping proteins, and virulence factors in bacteria. Accordingly, DAP epimerase is a promising antimicrobial target. Previous studies report that DAP epimerase exists as a monomeric enzyme. However, we show using analytical ultracentrifugation, X-ray crystallography, and enzyme kinetic analyses that DAP epimerase from Escherichia coli exists as a functional dimer in solution and the crystal state. Furthermore, the 2.0-Å X-ray crystal structure of the E. coli DAP epimerase dimer shows for the first time that the enzyme exists in an open, active conformation. The importance of dimerization was subsequently probed by using site-directed mutagenesis to generate a monomeric mutant (Y268A). Our studies show that Y268A is catalytically inactive, thus demonstrating that dimerization of DAP epimerase is essential for catalysis. Molecular dynamics simulations indicate that the DAP epimerase monomer is inherently more flexible than the dimer, suggesting that dimerization optimizes protein dynamics to support function. Our findings offer insight into the development of novel antimicrobial agents targeting the dimeric antibiotic target DAP epimerase.  相似文献   

2.
Diaminopimelate (DAP) epimerase is a key enzyme for the biosynthesis of lysine in plants. Lysine is an essential dietary nutrient for mammals. In both plants and bacteria, DAP epimerase catalyzes the interconversion of ll-DAP and dl(meso)-DAP. The absence of a mammalian homolog makes DAP epimerase a promising target for the design of novel herbicides and antibacterials. This enzyme requires no cofactors and it functions through an unusual mechanism involving two cysteine residues acting in concert and alternating as a base (thiolate) and as an acid (thiol). The present study reports the crystal structures of two enzyme-inhibitor complexes of DAP epimerase from Arabidopsis thaliana with different isomers of the irreversible inhibitor and substrate mimic, 2-(4-amino-4-carboxybutyl)-aziridine-2-carboxylate, at 1.95 and 2.3 Å resolution. These structures provide the first atomic details of a plant amino acid racemase. Structural analysis reveals that ligand binding to a cleft between the two domains of the enzyme is accompanied by domain closure with two strictly conserved cysteine residues, Cys99 and Cys254, optimally positioned to perform acid/base catalysis via a carbanion stabilization mechanism on the stereogenic α-carbon atom of the amino acid. Stereochemical control in catalysis is achieved by means of a highly symmetric catalytic site that can accommodate both the l and d stereogenic centers of DAP at the proximal site, whereas specific interactions at the distal site require only the l configuration. Structural comparisons of the plant enzyme with its bacterial counterpart from Haemophilus influenzae reveal significant conservation of amino acid residues around the active site that extends to their three-dimensional structures and catalytic mechanism.  相似文献   

3.
Analogs 1-8 of diaminopimelic acid (DAP) were synthesized and tested for inhibition of purified meso-DAP D-dehydrogenase from Bacillus sphaericus and of LL-DAP epimerase from Escherichia coli. The dehydrogenase was assayed by monitoring NADPH formation spectrophotometrically at 340 nm. N-Hydroxy DAP 4, N-amino DAP 5, and 4-methylene DAP 6 are substrates of the dehydrogenase with relative rates exceeding those of the meso isomers of the thia analogs 1ab, 2ab, and 3ab. DAP epimerase was assayed by coupling the epimerization of LL-DAP to DL-DAP (Km = 0.26 mM) with the dehydrogenase-catalyzed oxidation of DL-DAP by NADP. Lanthionine isomers 1ab and 1c were stronger inhibitors of the epimerase (Ki = 0.18 mM, Ki' = 0.67 mM, and Ki = 0.42 mM, respectively) than the corresponding meso-sulfoxide 2ab or the meso-sulfone 3ab. Other isomers of 2 and 3, as well as compounds 7 and 8, showed no epimerase inhibition. N-Hydroxy DAP 4 was the most potent competitive inhibitor (Ki = 0.0056 mM) of the epimerase, whereas N-amino DAP 5 is weaker (Ki = 2.9 mM) and 4-methylene DAP 6 is a noncompetitive inhibitor (Ki' = 0.95 mM). Although none of the analogs tested showed time-dependent inactivation of either enzyme, compounds 4, 5, 6, and 7 display substantial antibacterial activities. Possible mechanisms of epimerase inhibition and significance of the DAP pathway as a target for antibiotics are discussed.  相似文献   

4.
Diaminopimelate (DAP) epimerase catalyzes the stereoinversion of ll-DAP to meso-DAP, a precursor of l-lysine and an essential component of the bacterial peptidoglycan. This function is vital to bacteria and the enzyme therefore represents an attractive target for the design of novel anti-bacterials. DAP epimerase belongs to the group of PLP-independent amino acid racemases that function through a rather unusual mechanism involving two cysteines acting in concert as a base (thiolate) and an acid (thiol). We have solved the crystal structures of the apo-forms of DAP epimerase mutants (C73S and C217S) from Haemophilus influenzae at 2.3A and 2.2A resolution, respectively. These structures provide a snapshot of the enzyme in the first step of the catalytic cycle. Comparisons with the structures of the inhibitor-bound form reveal that the enzyme adopts an 'open conformation' in the absence of substrates or inhibitors with the two active site cysteines existing as a thiol-thiolate pair. Substrate binding to the C-terminal domain triggers the closure of the N-terminal domain coupled with tight encapsulation of the ligand, stabilization of the conformation of an active site loop containing Cys73 and expulsion of water molecules with concomitant desolvation of the thiolate base. This structural rearrangement is critical for catalysis.  相似文献   

5.
Extracts from Chlamydomonas, corn, soybean and tobacco were tested for enzymes of the lysine biosynthetic pathway. Dihydrodipicolinic acid (DHD) synthase, DHD reductase, diaminopimelate (DAP) epimerase and DAP decarboxylase were present in all. However, in contrast to the report of Wenko et al., meso-DAP dehydrogenase could not be detected in extracts prepared from soybean. Moreover, it was not found in Chlamydomonas, corn and tobacco as well. In order to set an upper limit to the amount of meso-DAP dehydrogenase that might be present, reconstruction experiments were performed with soybean and corn extracts in which the conversion of dihydrodipicolinate to lysine was made dependent on the addition of limited amounts of the meso-DAP dehydrogenase purified from Bacillus sphaericus. The presence of DAP epimerase and the absence of meso-DAP dehydrogenase indicates that the meso-DAP dehydrogenase abbreviated pathway for lysine synthesis is not operative in plants.  相似文献   

6.
Glycyrrhetic acid, derived from a main component of liquorice, was converted to 3-ketoglycyrrhetic acid reversibly by rat liver homogenates in the presence of NADPH or NADP+. Glycyrrhetic acid-oxidizing and 3-ketoglycyrrhetic acid-reducing activities were localized in microsomes among the subcellular fractions of rat liver. Glycyrrhetic acid-oxidizing activity and 3-ketoglycyrrhetic acid-reducing activities showed pH optima at 6.3 and 8.5, respectively, and required NADP+ or NAD+ and NADPH or NADH, respectively, indicating that these activities were due to glycyrrhetinate dehydrogenase. The dehydrogenase was not solubilized from the membranes by the treatment with 1 M NaCl or sonication, indicating that the enzyme is a membrane component. The dehydrogenase was solubilized with detergents such as Emalgen 913, Triton X-100 and sodium cholate, and then separated from 3β-hydroxysteroid dehydrogenase (5β-androstan-3β-ol-17-one-oxidizing activity) by butyl-Toyopearl 650 M column chromatography. Partially purified enzyme catalyzed the reversible reaction between glycyrrhetic acid and 3-ketoglycyrrhetic acid, but was inactive toward 3-epiglycyrrhetic acid and other steroids having the 3β-hydroxyl group. The enzyme required NADP+ and NADPH for the highest activities of oxidation and reduction, respectively, and NAD+ and NADH for considerable activities, similar to the results with microsomes. From these results the enzyme is defined as glycyrrhetinate dehydrogenase, being quite different from 3β-hydroxysteroid dehydrogenase of Ruminococcus sp. from human intestine, which is active for both glycyrrhetic acid and steroids having the 3β-hydroxyl group.  相似文献   

7.
A gene named ltsA was earlier identified in Rhodococcus and Corynebacterium species while screening for mutations leading to increased cell susceptibility to lysozyme. The encoded protein belonged to a huge family of glutamine amidotransferases whose members catalyze amide nitrogen transfer from glutamine to various specific acceptor substrates. We here describe detailed physiological and biochemical investigations demonstrating the specific role of LtsA protein from Corynebacterium glutamicum (LtsACg) in the modification by amidation of cell wall peptidoglycan diaminopimelic acid (DAP) residues. A morphologically altered but viable ΔltsA mutant was generated, which displays a high susceptibility to lysozyme and β-lactam antibiotics. Analysis of its peptidoglycan structure revealed a total loss of DAP amidation, a modification that was found in 80% of DAP residues in the wild-type polymer. The cell peptidoglycan content and cross-linking were otherwise not modified in the mutant. Heterologous expression of LtsACg in Escherichia coli yielded a massive and toxic incorporation of amidated DAP into the peptidoglycan that ultimately led to cell lysis. In vitro assays confirmed the amidotransferase activity of LtsACg and showed that this enzyme used the peptidoglycan lipid intermediates I and II but not, or only marginally, the UDP-MurNAc pentapeptide nucleotide precursor as acceptor substrates. As is generally the case for glutamine amidotransferases, either glutamine or NH4+ could serve as the donor substrate for LtsACg. The enzyme did not amidate tripeptide- and tetrapeptide-truncated versions of lipid I, indicating a strict specificity for a pentapeptide chain length.  相似文献   

8.
NADP-malate dehydrogenase was purified from leaves of Zea mays in the absence of thiol-reducing agents by (NH4)2SO4, polyethylene glycol, and pH fractionation followed by dye-ligand affinity chromatography and gel filtration. The purified enzyme is completely inactive (no activity detected between pH 6 and 9) but can be reactivated by thiol-reducing agents including dithiothreitol and thioredoxin. The active enzyme shows distinctly alkaline pH optima when assayed in either direction; Km values at pH 8.5 are oxaloacetate, 18 μm; malate, 24 mm; NADPH, 50 μm; and NADP, 45 μm. The reduction of oxaloacetate is inhibited by NADP (competitive with respect to NADPH, Ki = 50 μm). The molecular weight of the native inactive or active enzyme is 150,000 with subunits of Mr 38,000. Active enzyme is much more sensitive (>50-fold) to heat denaturation than is the inactive enzyme and is irreversibly inactivated by N-ethylmaleimide whereas the inactive enzyme is insensitive to this reagent. The active and inactive forms of NADP-malate dehydrogenase are assumed to correspond to dithiol and disulfide forms of the enzyme, respectively. The relative coenzyme-binding affinities of inactive NADP-malate dehydrogenase differ by a factor of 102 from the binding affinities for active NADP-malate dehydrogenase and 104 for non-thiol-regulated NAD-specific malate dehydrogenase. It is proposed that the 100-fold change in differential binding of NADP and NADPH upon conversion of NADP-malate dehydrogenase to the disulfide form may sufficiently alter the equilibrium of the central enzyme-substrate complexes, and hence the catalytic efficiency of the enzyme, to explain the associated loss of activity.  相似文献   

9.
Equilibrium dialysis indicates that rat liver glucose-6-P dehydrogenase binds two molecules of NADP+ per subunit with a dissociation constant of 0.6 × 10?6 M. The NADP+ free enzyme will not bind glucose-6-P indicating a compulsory order of substrate binding. Development of an isotopic assay allowed a direct measurement of the effect of physiological alterations in the NADP+/NADPH ratio on the activity of glucose-6-P and 6-phosphogluconate dehydrogenases. A combination of enzyme induction and altered NADP+/NADPH ratios could produce 30–50 fold changes in the capacity of these enzymes to produce NADPH during alterations in the nutritional state of the animal.  相似文献   

10.
Pyridine nucleotide specificity of barley nitrate reductase   总被引:6,自引:4,他引:2       下载免费PDF全文
Dailey FA  Kuo T  Warner RL 《Plant physiology》1982,69(5):1196-1199
NADPH nitrate reductase activity in higher plants has been attributed to the presence of NAD(P)H bispecific nitrate reductases and to the presence of phosphatases capable of hydrolyzing NADPH to NADH. To determine which of these conditions exist in barley (Hordeum vulgare L. cv. Steptoe), we characterized the NADH and NADPH nitrate reductase activities in crude and affinity-chromatography-purified enzyme preparations. The pH optima were 7.5 for NADH and 6 to 6.5 for the NADPH nitrate reductase activities. The ratio of NADPH to NADH nitrate reductase activities was much greater in crude extracts than it was in a purified enzyme preparation. However, this difference was eliminated when the NADPH assays were conducted in the presence of lactate dehydrogenase and pyruvate to eliminate NADH competitively. The addition of lactate dehydrogenase and pyruvate to NADPH nitrate reductase assay media eliminated 80 to 95% of the NADPH nitrate reductase activity in crude extracts. These results suggest that a substantial portion of the NADPH nitrate reductase activity in barley crude extracts results from enzyme(s) capable of converting NADPH to NADH. This conversion may be due to a phosphatase, since phosphate and fluoride inhibited NADPH nitrate reductase activity to a greater extent than the NADH activity. The NADPH activity of the purified nitrate reductase appears to be an inherent property of the barley enzyme, because it was not affected by lactate dehydrogenase and pyruvate. Furthermore, inorganic phosphate did not accumulate in the assay media, indicating that NADPH was not converted to NADH. The wild type barley nitrate reductase is a NADH-specific enzyme with a slight capacity to use NADPH.  相似文献   

11.
ω-Hydroxyfatty acid dehydrogenase (ω-hydroxyfatty acid:NADP oxidoreductase) catalyzes the reaction ω-hydroxyfatty acid + NADP ? ω-oxofatty acid + NADPH +H+. In wound-healing potato tuber disks, the ω-oxofatty acid generated by this enzyme is further oxidized to the corresponding dicarboxylic acid by a separate enzyme, ω-oxofatty acid dehydrogenase. ω-Hydroxy acid dehydrogenase, but not ω-oxo acid dehydrogenase, was found to be induced by wounding potato tubers. ω-Hydroxy acid dehydrogenase has been purified 600-fold to near homogeneity from wound-healing potato tuber disks by a combination of gel filtration, anion-exchange, and hydroxylapatite chromatography followed by NADP-Sepharose affinity chromatography, in about 1% yield. The molecular weight and Stokes radius of this enzyme as determined by gel exclusion chromatography are 60,000 and 31 Å, respectively. Sodium dodecyl sulfate-gel electrophoresis gave a molecular weight of 31,000, indicating that the deydrogenase is a dimer with subunits of similar molecular weight. The pH optima for the reaction in the forward and reverse directions are 9.5 and 8.5, respectively, and V in the forward and reverse directions are 140 and 3200 nmol/min/mg, respectively. Apparent Km values for NADP, 16-hydroxyhexadecanoic acid, NADPH, and 16-oxohexadecanoic acid are 100, 20, 5, and 7 μm respectively. The equilibrium constant of the reaction at pH 9.5 and 30 °C is 1.4 × 10?9m. The enzyme preparation did not show any stereospecificity for hydride transfer from NADPH to 16-oxohexadecanoic acid.  相似文献   

12.
Mitochondrial NADH dehydrogenase has been purified to homogeneity by resolution of Complex I from beef heart mitochondria with the chaotrope NaClO4 and precipitation of the enzyme with ammonium sulfate. The enzyme is water-soluble, has a molecular weight of 69,000 ± 1000 as determined by gel filtration on Sephadex G-100 and agarose 1.5 M. It is an iron-sulfur flavoprotein, with the ratio of flavin (FMN) to nonheme iron to labile sulfide being 1:5–6:5–6. The FMN content suggests a minimum molecular weight of 74,000 ± 3000 for the enzyme. NADH dehydrogenase is composed of three subunits with apparent Mr values, as determined by acrylamide gel electrophoresis as well as by gel filtration on agarose 5 M both in the presence of sodium dodecyl sulfate, of about 51,000, 24,000, and 9–10,000. Coomassie blue stain intensities of the subunits on acrylamide gels suggest that they are present in NADH dehydrogenase in equimolar amounts. However, summation of the apparent Mr values of the dodecyl sulfate-treated subunits appears to overestimate the molecular weight of the native enzyme. The amino acid compositions of NADH dehydrogenase and of each of the isolated and purified subunits have been determined. NADH dehydrogenase catalyzes the oxidation of NADH and NADPH by quinones, ferric compounds, and NAD (3-acetylpyridine adenine dinucleotide was used). All the activities of NADH dehydrogenase are greatly stimulated by addition of guanidine (up to 150 mm), alkylguanidines, arginine, and arginine methyl ester to the assay medium. Phosphoarginine had no effect. These results pointed to the importance of the positively charged guanido group, which appears to interact with and neutralize the negative charges on NAD(P)H and thereby allow for better enzyme-substrate interaction. In the absence of guanidine, NADPH is essentially unoxidized by the enzyme at pH values above 6.0. However, both NADPH dehydrogenase and NADPH → NAD transhydrogenase activities increase dramatically as the assay pH is lowered below pH = 6. Since the pK of the 2′-phosphate of NADPH is 6.1, it appears that the above pH effect is related to protonation of the 2′-phosphate, thus rendering NADPH a closer electronic analog of NADH, which is the primary substrate of the enzyme.  相似文献   

13.
14.
NAD(P)H dehydrogenase was purified approximately 480-fold from Saccharomyces cerevisiae with 6.5% activity yield. The enzyme was homogeneous on polyacrylamide gel electrophoresis. The molecular weight of the enzyme was estimated to be 40,000–44,000 by gel filtration on Sephadex G-150 column chromatography and SDS-polyacrylamide gel electrophoresis. The Km values for NADPH and NADH were 7.3 μM and 0.1 mM, respectively. The activity of the enzyme increased approximately 4-fold with Cu2+. FAD, FMN and cytochrome c were not effective as electron acceptors, although Fe(CN)63− was slightly effective. NADH generated by the reaction of lactaldehyde dehydrogenase in the glycolytic methylglyoxal pathway will be reoxidized by NAD(P)H dehydrogenase. NAD(P)H dehydrogenase thus may contribute to the reduction/oxidation system in the glycolytic methylglyoxal pathway to maintain the flux of methylglyoxal to lactic acid via lactaldehyde.  相似文献   

15.
6-Phosphogluconate dehydrogenase (6PGDH), the third enzyme of the pentose phosphate pathway (PPP), is essential for biosyntheses and oxidative stress defence. It also has the function of depleting 6PG, whose accumulation induces cell senescence. 6PGDH is a proposed drug target for African trypanosomiasis caused by Trypanosoma brucei and for other microbial infections and cancer. Gel filtration, density gradient sedimentation, cross-linking and dynamic light scattering were used to assay the oligomerization state of T. brucei 6PGDH in the absence and presence of several ligands. The enzyme displays a dimer–tetramer equilibrium and NADPH (but not NADP) reduces the rate of approach to equilibrium, while 6PG is able to antagonize the NADPH effect. The different behaviour of the two forms of coenzyme appears to be related to the differences in ΔCp, with NADP binding ΔCp closer to what is expected of crystallographic structures, while NADPH ΔCp is three times larger. The estimated dimer–tetramer association constant is 1.5 · 106 M? 1, and the specific activity of the tetramer is about 3 fold higher than the specific activity of the dimer. Thus, cellular conditions promoting tetramer formation could allow an efficient clearing of 6PG. Experiments carried out on sheep liver 6PGDH indicate that tetramerization is a specificity of the parasite enzyme.  相似文献   

16.
An improved light-dependent assay was used to characterize the NAD(P)H dehydrogenase (NDH) in thylakoids of barley (Hordeum vulgare L.). The enzyme was sensitive to rotenone, confirming the involvement of a complex I-type enzyme. NADPH and NADH were equally good substrates for the dehydrogenase. Maximum rates of activity were 10 to 19 μmol electrons mg−1 chlorophyll h−1, corresponding to about 3% of linear electron-transport rates, or to about 40% of ferredoxin-dependent cyclic electron-transport rates. The NDH was activated by light treatment. After photoactivation, a subsequent light-independent period of about 1 h was required for maximum activation. The NDH could also be activated by incubation of the thylakoids in low-ionic-strength buffer. The kinetics, substrate specificity, and inhibitor profiles were essentially the same for both induction strategies. The possible involvement of ferredoxin:NADP+ oxidoreductase (FNR) in the NDH activity could be excluded based on the lack of preference for NADPH over NADH. Furthermore, thenoyltrifluoroacetone inhibited the diaphorase activity of FNR but not the NDH activity. These results also lead to the conclusion that direct reduction of plastoquinone by FNR is negligible.  相似文献   

17.
All the glutamate dehydrogenase activity in developing castor bean endosperm is shown to be located in the mitochondria. The enzyme can not be detected in the plastids, and this is probably not due to the inactivation of an unstable enzyme, since a stable enzyme can be isolated from castor bean leaf chloroplasts. The endosperm mitochondrial glutamate dehydrogenase consists of a series of differently charged forms which stain on polyacrylamide gel electrophoresis with both NAD+ and NADP+. The chloroplast and root enzymes differ from the endosperm enzyme on polyacrylamide gel electrophoresis. The amination reaction of all the enzymes is affected by high salt concentrations. For the endosperm enzyme, the ratio of activity with NADH to that with NADPH is 6.3 at 250 millimolar NH4Cl and 1.5 at 12.5 millimolar NH4Cl. Km values for NH4+ and NAD(P)H are reduced at low salt concentrations. The low Km values for the nucleotides may favor a role for glutamate dehydrogenase in ammonia assimilation in some situations.  相似文献   

18.
Activity analyses of pure dihydrofolate reductase from amethopterin-resistant Lactobacillus casei conducted with commercial sources of NADPH yielded a progression of nonlinear assay tracings whose shapes were both pH dependent and reminiscent of classical product inhibition. The extent of curving of the assay tracings was dependent on the source and age of the commercial NADPH and was enhanced as the pH was decreased from 7.5 to 5.0. Under these conditions a “pseudo”-pH-activity profile, exhibiting a maximal specific activity of 9 units/mg of protein between pH 7.0 and 7.5, was found. In contrast, freshly prepared NADPH provided strictly linear assay tracings over the pH range of 8.5 to 5.0, yielding uniformly higher specific activities than those observed with commercial NADPH. The new pH-activity profile was characterized by a broad optimum between pH 5.0 and 6.0, with a maximal specificity activity of 24.9 units/ mg in 0.1m potassium phosphate in the absence of added salt. The curving phenomenon and pseudo-pH optimum observed with commercial NADPH is attributed to the presence of minor but potent inhibitory impurities in these coenzyme preparations. Optimal concentrations of monovalent (~0.1 m) and divalent (~0.05 m) salts activated the enzyme between 1.5- and 1.7-fold, resulting in maximal specific activities in the range of 34 to 39 units/mg. A similar extent of activation was observed in 0.8 m Tris-acetate buffer, pH 5.5. At concentrations of monovalent salts above 0.5 m and of divalent salts above 0.2 m a reduction in salt-dependent activation and, in some cases, inhibition of activity were obtained. Substrate specificity studies indicated that the V for folate at saturating levels is 1% of that for dihydrofolate. Deamino-NADPH yielded V values 1.4-fold higher than that for NADPH, while acetylpyridine-NADPH and thio-NADPH provided values 6.5- and 235-fold lower, respectively, than the value with the natural coenzyme. Gel electrophoresis studies reflected a similar trend of selectivity in the interaction of NADPH and its analogs to form stable binary complexes. Stable ternary complexes of enzyme and amethopterin were formed with NADPH, deamino-NADPH, thio-NADPH, and acetylpyridine-NADPH. Although neither dihydrofolate nor NADP+ and its analog form stable complexes with L. casei dihydrofolate reductase, both NADP+ and deamino-NADP+ interact with enzyme and dihydrofolate to generate stable ternary complexes.  相似文献   

19.
Methylated inositol, d ‐pinitol (3‐O‐methyl‐d ‐chiro‐inositol), is a common constituent in legumes. It is synthesized from myo‐inositol in two reactions: the first reaction, catalyzed by myo‐inositol‐O‐methyltransferase (IMT), consists of a transfer of a methyl group from S‐adenosylmethionine to myo‐inositol with the formation of d ‐ononitol, while the second reaction, catalyzed by d ‐ononitol epimerase (OEP), involves epimerization of d ‐ononitol to d ‐pinitol. To identify the genes involved in d ‐pinitol biosynthesis in a model legume Medicago truncatula, we conducted a BLAST search on its genome using soybean IMT cDNA as a query and found putative IMT (MtIMT) gene. Subsequent co‐expression analysis performed on publicly available microarray data revealed two potential OEP genes: MtOEPA, encoding an aldo‐keto reductase and MtOEPB, encoding a short‐chain dehydrogenase. cDNAs of all three genes were cloned and expressed as recombinant proteins in E. coli. In vitro assays confirmed that putative MtIMT enzyme catalyzes methylation of myo‐inositol to d ‐ononitol and showed that MtOEPA enzyme has NAD+‐dependent d ‐ononitol dehydrogenase activity, while MtOEPB enzyme has NADP+‐dependent d ‐pinitol dehydrogenase activity. Both enzymes are required for epimerization of d ‐ononitol to d ‐pinitol, which occurs in the presence of NAD+ and NADPH. Introduction of MtIMT, MtOEPA, and MtOEPB genes into tobacco plants resulted in production of d ‐ononitol and d ‐pinitol in transformants. As this two‐step pathway of d ‐ononitol epimerization is coupled with a transfer of reducing equivalents from NADPH to NAD+, we speculate that one of the functions of this pathway might be regeneration of NADP+ during drought stress.  相似文献   

20.
Mannitol dehydrogenase (mannitol: NADP+ 2-oxidoreductase: EC 1.1.1.138) was isolated from Agaricus bisporus by fractionation with protamine sulphate and (NH4)2SO4, followed by chromatography on DEAE-Sephadex, then by affinity and gel chromatography. The products of enzyme reaction were identified by GLC and TLC. Km, optimum pH, MW and pI of the enzyme as well as the influence of temperature, ions and inhibitors on enzymic activity were determined. In the sugar reducing reaction, the enzyme was specific for fructose but, in the reverse direction, some structurally related polyols could substitute for mannitol. The enzyme was very sensitive to alterations in the NADP+/NADPH ratio. The results are discussed in relation to the possible role of mannitol dehydrogenase in fungal metabolism.  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