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.
Zhang GL  Wang CW  Li C 《Biotechnology letters》2012,34(8):1519-1523
The budC gene encoding the meso-2,3-BDH from Klebsiella pneumoniae XJ-Li was expressed in E. coli BL21 (DE3) pLys. Hypothetical amino acid sequence alignments revealed that the enzyme belongs to the short chain dehydrogenase/reductase family. After purification and refolding, the recombinant enzyme had activities of 218 U/mg for reduction of acetoin and 66 U/mg for oxidation of meso-2,3-butanediol. Highest activities were at pH 8.0 and 9.0 respectively. These are higher than other meso-2,3-butanediol dehydrogenases from K. pneumoniae. The low K (m) value (0.65 mM) for acetoin indicated that the enzyme can easily reduce acetoin to meso-2,3-butanediol. There were no significant activities towards 2R,3R-2,3-butanediol, 1,4-butanediol and 2S,3S-2,3-butanediol, suggesting that the enzyme has a high stereospecificity for the meso-dihydric alcohol.  相似文献   

2.
A (2R,3R)-2,3-butanediol dehydrogenase (BDH99::67) from Paenibacillus polymyxa ATCC 12321 was functionally characterized. The genetic characteristics of BDH99::67 are completely different from those of meso- and (2S,3S)-2,3-butanediol dehydrogenases. The results showed that BDH99::67 belongs to the medium-chain dehydrogenase/reductase superfamily and not to the short-chain dehydrogenase/reductase superfamily, to which meso- and (2S,3S)-2,3-butanediol dehydrogenases belong.  相似文献   

3.
A number of problems present themselves during the gas chromatographic-mass spectrometric assay of R,S-1,3-butanediol as its bis-tert-butyldimethylsilyl ether. To circumvent these problems, three labeled internal standards were synthesized: (i) R,S-1,3-[3,4-13C2]-butanediol, (ii) R,S-1,3-[1,1,3-2H3]butanediol, and (iii) R,S-1,3-[1,1,3-2H3,3,4-13C2]butanediol. The availability of internal standards with different degrees of labeling allows (i) assaying of either unlabeled or 13C-labeled R,S-1,3-butanediol and (ii) analysis of 1,3-butanediol in either blood or urine samples. Reproducible standard curves were obtained using both electron impact and ammonia chemical ionization modes. The latter provides greater sensitivity and a lower limit of detection (5 microM). We have also designed an indirect assay of S-3-hydroxybutyrate, a catabolite of R,S-1,3-butanediol, which is difficult to analyze by conventional methods. This assay relies on the difference between (i) the concentration of R,S-3-hydroxybutyrate assayed by gas chromatography-mass spectrometry and (ii) the concentration of R-3-hydroxybutyrate assayed enzymatically.  相似文献   

4.
Using a conserved sequence motif, a new gene (YAL060W) of the MDR family has been identified in Saccharomyces cerevisiae. The expressed protein was a stereoespecific (2R,3R)-2,3-butanediol dehydrogenase (BDH). The best substrates were (2R,3R)-2,3-butanediol for the oxidation and (3R/3S)-acetoin and 1-hydroxy-2-propanone for the reduction reactions. The enzyme is extremely specific for NAD(H) as cofactor, probably because the presence of Glu223 in the cofactor binding site, instead of the highly conserved Asp223. BDH is inhibited competitively by 4-methylpyrazole with a K(i) of 34 microM. Yeast could grow on 2,3-butanediol or acetoin as a sole energy and carbon sources, and a 3.6-fold increase in BDH activity was observed when cells were grown in 2,3-butanediol, suggesting a role of the enzyme in 2,3-butanediol metabolism. However, the disruption of the YAL060W gene was not lethal for the yeast under laboratory conditions, and the disrupted strain could also grow in 2,3-butanediol and acetoin. This suggests that other enzymes, in addition to BDH, can also metabolize 2,3-butanediol in yeast.  相似文献   

5.
The present work aims to block 2,3-butanediol synthesis in acetoin fermentation of Bacillus subtilis. First, we constructed a recombinant strain BS168D by deleting the 2,3-butanediol dehydrogenase gene bdhA of the B. subtilis168, and there was almost no 2,3-butanediol production in 20?g/L of glucose media. The acetoin yield of BS168D reached 6.61?g/L, which was about 1.5 times higher than that of the control B. subtilis168 (4.47?g/L). Then, when the glucose concentration was increased to 100?g/L, the acetoin yield reached 24.6?g/L, but 2.4?g/L of 2,3-butanediol was detected at the end of fermentation. The analysis of 2,3-butanediol chiral structure indicated that the main 2,3-butanediol production of BS168D was meso-2,3-butanediol, and the bdhA gene was only responsible for (2R,3R)-2,3-butanediol synthesis. Therefore, we speculated that there may exit another pathway relating to the meso-2,3-butanediol synthesis in the B. subtilis. In addition, the results of low oxygen condition fermentation showed that deletion of bdhA gene successfully blocked the reversible transformation between acetoin and 2,3-butanediol and eliminated the effect of dissolved oxygen on the transformation.  相似文献   

6.
Dimethylamine [DMA, (CH(3))(2)NH)] is abundantly present in human urine. Main sources of urinary DMA have been reported to include trimethylamine N-oxide, a common food component, and asymmetric dimethylarginine (ADMA), an endogenous inhibitor of nitric oxide (NO) synthesis. ADMA is excreted in the urine in part unmetabolized and in part after hydrolysis to DMA by dimethylarginine dimethylaminohydrolase (DDAH). Here we describe a GC-MS method for the accurate and rapid quantification of DMA in human urine. The method involves use of (CD(3))(2)NH as internal standard, simultaneous derivatization with pentafluorobenzoyl chloride and extraction in toluene, and selected-ion monitoring of m/z 239 for DMA and m/z 245 for (CD(3))(2)NH in the electron ionization mode. GC-MS analysis of urine samples from 10 healthy volunteers revealed a DMA concentration of 264+/-173 microM equivalent to 10.1+/-1.64 micromol/mmol creatinine. GC-tandem MS analysis of the same urine samples revealed an ADMA concentration of 27.3+/-15.3 microM corresponding to 1.35+/-1.2 micromol/mmol creatinine. In these volunteers, a positive correlation (R=0.83919, P=0.0024) was found between urinary DMA and ADMA, with the DMA/ADMA molar ratio being 10.8+/-6.2. Elevated excretion rates of DMA (52.9+/-18.5 micromol/mmol creatinine) and ADMA (3.85+/-1.65 micromol/mmol creatinine) were found by the method in 49 patients suffering from coronary artery disease, with the DMA/ADMA molar ratio also being elevated (16.8+/-12.8). In 12 patients suffering from end-stage liver disease, excretion rates of DMA (47.8+/-19.7 micromol/mmol creatinine) and ADMA (5.6+/-1.5 micromol/mmol creatinine) were found to be elevated, with the DMA/ADMA molar ratio (9.17+/-4.2) being insignificantly lower (P=0.46). Between urinary DMA and ADMA there was a positive correlation (R=0.6655, P<0.0001) in coronary artery disease, but no correlation (R=0.27339) was found in end-stage liver disease.  相似文献   

7.
Liu Z  Qin J  Gao C  Hua D  Ma C  Li L  Wang Y  Xu P 《Bioresource technology》2011,102(22):10741-10744
Production of highly pure (2S,3S)-2,3-butanediol ((2S,3S)-2,3-BD) and (3S)-acetoin ((3S)-AC) in high concentrations is desirable but difficult to achieve. In the present study, glucose was first transformed to a mixture of (2S,3S)-2,3-BD and meso-2,3-BD by resting cells of Klebsiella pneumoniae CICC 10011, followed by biocatalytic resolution of the mixture by resting cells of Bacillus subtilis 168. meso-2,3-BD was transformed to (3S)-AC, leaving (2S,3S)-2,3-BD in the reaction medium. Using this approach, 12.5 g l(-1) (2S,3S)-2,3-BD and 56.7 g l(-1) (3S)-AC were produced. Stereoisomeric purity of (2S,3S)-2,3-BD and enantiomeric excess of (3S)-AC was 96.9 and 96.2%, respectively.  相似文献   

8.
1. R-Acetate was generated enzymically from R-acetate in the sequence acetate leads to malate leads to oxaloacetate leads to acetate, and S-acetate likewise from S-acetate. It was concluded that the formation of malate on malate synthase involves the operation of a normal isotopic effect combined with inversion of configuration. The malate synthase kH/k2H was determined as 3.7 +/- 0.5 by a method which yields results independently of the stereochemical purity of the chiral acetates used initially. 2. R-Acetate was also generated from R-acetate in the sequence acetate leads to citrate leads to malate leads to oxaloacetate leads to acetate, and S-acetate likewise from S-acetate. The conclusion is the same as given above, but refers to the formation of citrate on the re-synthase. 3. 2S,3R-[2-2H1,3-2H1,3H1]Malate and 2S,3S-[2-2H1,3-2H1]malate were prepared from 2S-[2,3-2H3]malate by treatment with fumarase in tritiated water and normal water, respectively. It was assumed that these malate specimens were pure with respect to chirality as generated by isotopic labelling. 4. These two malate specimens were partially converted (about 9%) to acetates in conditions where no racemization at the level of transiently formed oxaloacetate occurred. That no racemization took place was demonstrated experimentally. Oxidative enzymic hydrolysis of 2S,3R-[2-2H1,3-2H1,3H1]malate in normal water and of 2S,3S-[2-2H1,3-2H1]malate in tritiated water produced S-[2H1,3H1]acetate and R-[2H1,3H1]acetate, respectively. 5. The isolated R-[2H1,3H1]acetate and S-[2H1,3H1]acetate on configurational analysis yielded malates which in the presence of fumarase retained 79.7 +/- 0.7% and 20.3 +/- 0.9%, respectively, of their total tritium content. The symmetric deviation from the 50% value found with [3H1]acetate strengthens the conclusion that stereochemically pure chiral acetates were analyzed. The malate synthase kH/k2H was determined from the data of this study as 3.9 +/- 0.2. 6. The average of the values given under paragraphs 1 and 5 for the isotopic discrimination on malate synthase corresponds to kH/k2H=3.8 +/- 0.1. It was concluded that the configurational analysis of stereochemically pure R-[2H1,3H1]acetate and S-[2H1,3H1]acetate yields malates which in the presence of fumarase retain 79 +/- 2% and 21 +/- 2%, respectively, of their total tritium content. Hence, a deviation of 29 +/- 2% from the 50% value represents the actual amplitude of the configurational assay. 7. Outlines are given for an enzymic generation of chiral acetates in preparative scale.  相似文献   

9.
2,3-butanediol dehydrogenase (BDH, EC 1.1.1.76) also known as acetoin reductase (AR, EC 1.1.1.4) is the key enzyme converting acetoin (AC) into 2,3-butanediol (BD) and undertaking the irreversible conversion of diacetyl to acetoin in various microorganisms. The existence of three BDHs (R,R-, meso-, and S,S-BDH) product different BD isomers. Catalyzing mechanisms of meso- and S,S-BDH have been understood with the assistance of their X-ray crystal structures. However, the lack of structural data for R,R-BDH restricts the integral understanding of the catalytic mechanism of BDHs. In this study, we successfully crystallized and solved the X-ray crystal structure of Bacillus subtilis R,R-BDH. A zinc ion was found locating in the catalytic center and coordinated by Cys37, His70 and Glu152, helping to stabilize the chiral substrates observed in the predicted molecular docking model. The interaction patterns of different chiral substrates in the molecular docking model explained the react priority measured by the enzyme activity assay of R,R-BDH. Site-directed mutation experiments determined that the amino acids Cys37, Thr244, Ile268 and Lys340 are important in the catalytically active center. The structural information of R,R-BDH presented in this study accomplished the understanding of BDHs catalytic mechanism and more importantly provides useful guidance for the directional engineering of R,R-BDH to obtain high-purity monochiral BD and AC.  相似文献   

10.
1. In addition to the known 2R,3R- and 2R, 3S-2,3-dihydroxy-3-methylpentanoic acids (DHI), the 1S,3S- and sS,DR-isomers were prepared. 2S-2,3-Dihydroxy-3-methylbutanoic acid (DHV) was also prepared in addition to the known 2R-isomer. 2. The six dihydroxy acids were examined for their ability to promote the growth of isoleucine-valine (ilv)-requiring strains of Salmonella typhimurium and to serve as substrates for the alpha,beta-dihydroxyacid dehydratase of the same organism. 3. Only 2R,3R-2,3-dihydroxy-3-methylpentanoic and 2R-2,3-dihydroxy-3-methylbutanoic acids supported growth of the ilv strains of S. typhimurium. 4. alpha,beta-Dihydroxyacid dehydratase utilized the three isomers with the 2R-configuration as substrates but not those with the 2S-configuration. 5. In an additional growth study that utilized the 3R- and 3S-isomers of 3-methyl-2-oxopentanoic acid, the alpha-keto acid analogue of isoleucine, only the 3S-isomer supported growth. 6. It is concluded that the mechanism of action of the dehydratase is stereospecific in that the proton that is attached to C-3 of the substrate occupies the same steriochemical position as the departing hydroxyl group (Fig. 6).  相似文献   

11.
The completion of the Saccharomyces cerevisiae genome project in 1996 showed that almost 60% of the potential open reading frames of the genome had no experimentally determined function. Using a conserved sequence motif present in the zinc-containing medium-chain alcohol dehydrogenases, we found several potential alcohol dehydrogenase genes with no defined function. One of these, YAL060W, was overexpressed using a multicopy inducible vector, and its protein product was purified to homogeneity. The enzyme was found to be a homodimer that, in the presence of NAD(+), but not of NADP, could catalyze the stereospecific oxidation of (2R,3R)-2, 3-butanediol (K(m) = 14 mm, k(cat) = 78,000 min(-)(1)) and meso-butanediol (K(m) = 65 mm, k(cat) = 46,000 min(-)(1)) to (3R)-acetoin and (3S)-acetoin, respectively. It was unable, however, to further oxidize these acetoins to diacetyl. In the presence of NADH, it could catalyze the stereospecific reduction of racemic acetoin ((3R/3S)- acetoin; K(m) = 4.5 mm, k(cat) = 98,000 min(-)(1)) to (2R,3R)-2,3-butanediol and meso-butanediol, respectively. The substrate stereospecificity was determined by analysis of products by gas-liquid chromatography. The YAL060W gene product can therefore be classified as an NAD-dependent (2R,3R)-2,3-butanediol dehydrogenase (BDH). S. cerevisiae could grow on 2,3-butanediol as the sole carbon and energy source. Under these conditions, a 3. 5-fold increase in (2R,3R)-2,3-butanediol dehydrogenase activity was observed in the total cell extracts. The isoelectric focusing pattern of the induced enzyme coincided with that of the pure BDH (pI 6.9). The disruption of the YAL060W gene was not lethal for the yeast under laboratory conditions. The disrupted strain could also grow on 2,3-butanediol, although attaining a lesser cell density than the wild-type strain. Taking into consideration the substrate specificity of the YAL060W gene product, we propose the name of BDH for this gene. The corresponding enzyme is the first eukaryotic (2R, 3R)-2,3-butanediol dehydrogenase characterized of the medium-chain dehydrogenase/reductase family.  相似文献   

12.
An Na+-stimulated Mg2+-transport system in human red blood cells   总被引:5,自引:0,他引:5  
The initial rate of net Mg2+ efflux was measured in human red blood cells by atomic absorption. In fresh erythrocytes incubated in Na+,K+-Ringer's medium this rate was 7.3 +/- 2.8 mumol/l cells per h (mean +/- S.D. of 14 subjects) with an energy of activation of 13 200 cal/mol. Cells with total Mg2+ contents ([ Mg]i) ranging from 1.8 to 24 mmol/l cells were prepared by using a modified p-chloromercuribenzenesulphonate method. Mg2+ efflux was strongly stimulated by increases in [Mg]i and in external Na+ concentrations ([ Na]o). A kinetic analysis of Mg2+ efflux as a function of [Mg]i and [Na]o revealed the existence of two components: an Na+-stimulated Mg2+ efflux, which exhibited a Michaelian-like dependence of free internal Mg2+ content (apparent dissociation constant = 2.6 +/- 1.4 mmol/l cells; mean +/- S.D. of six subjects) and on external Na+ concentration (apparent dissociation constant = 20.5 +/- 1.9 mM; mean +/- S.D. of four subjects) and a variable maximal rate ranging from 35 to 370 mumol/l cells per h, and an Na+-independent Mg2+ efflux, which showed a linear dependence on internal Mg2+ content with a rate constant of (6.6 +/- 0.7) X 10(-3) h-1. Fluxes catalyzed by the Na+-stimulated Mg2+ carrier were partially dependent on the ATP content of the cells and completely inhibited by quinidine (IC50 = 50 microM) and by Mn2+ (IC50 = 0.5-1.0 mM).  相似文献   

13.
An enantioselective assay for S-(-)- and R-(+)-propranolol in transgenic Chinese hamster CHL cell lines, expressing human cytochrome P450 (CYP), was developed. The method involves extraction of propranolol from the S(9) incubates, using S-(+)-propafenone as internal standard, chiral derivatization with 2,3,4,6-tetra-O-beta-D-glucopranosyl isothiocyanate and quantitation by reversed phase high-performance liquid chromatography system with UV detection (lambda=220 nm). A baseline separation of propranolol enantiomers was achieved on a 5-microm reverse-phase ODS column, with a mixture of methanol/water/glacial acetic acid (67:33:0.05, v/v) as mobile phase. The assay is linear from 5 to 500 microM for each enantiomer. The analytical method affords average recoveries of 99.2% and 98.8% for S-(-)- and R-(+)-propranolol, respectively. The limit of quantitation for the method is 5 microM for both S-(-)- and R-(+)-propranolol. The reproducibility of the assay is satisfactory (RSD < 10%). The method allowed study of the depletion of S-(-)- and R-(+)-propranolol in transgenic Chinese hamster CHL cell lines expressing CYP3A4, CYP2C18 and CYP2C9.  相似文献   

14.
New analogues of 3beta-hydroxy-5alpha-cholest-8(14)-en-15-one (15-ketosterol) with modified 17-chains [(22S,23S,24S)- and (22R,23R,24S)-3beta-hydroxy-24-methyl-22,23-oxido-5alpha-cholest-8(14)-en-15-ones and (22RS,23xi,24S)-24-methyl-5alpha-cholesta-3beta,22,23-triol-15-one] were synthesized from (22E,24S)-3beta-acetoxy-24-methyl-5alpha-cholesta-8(14),22-dien-15-one. The chiralities of their 22 and 23 centers were determined by NMR spectroscopy. The isomeric 22,23-epoxides effectively inhibited cholesterol biosynthesis in hepatoma Hep G2 cells (IC50 0.9 +/- 0.2 and 0.7 +/- 0.2 microM, respectively), and their activities significantly exceeded those of 15-ketosterol (IC50 4.0 +/- 0.5 microM), (22E,24S)-3beta-hydroxy-24-methyl-5alpha-cholesta-8(14),22-dien-15-one (IC50 3.1 +/- 0.4 microM), and the 3beta,22,23-triol synthesized (IC50 6.0 +/- 1.0 microM). The English version of the paper: Russian Journal of Bioorganic Chemistry, 2005, vol. 31, no. 3; see also http://www.maik.ru.  相似文献   

15.
16.
Glutathione (GSH) conjugation of the chiral compound 2-bromo-3-phenylpropionic acid (BPP) was studied in vitro and in the rat in vivo. GSH conjugation of BPP, catalyzed by a mixture of glutathione-S-transferases (GST's) from rat liver cytosol in vitro, was stereoselective: at a substrate concentration of 250 microM, (R)-BPP was more rapidly conjugated than (S)-BPP (R/S-ratio = 2.6). The blood elimination kinetics of the separate BPP enantiomers and the biliary excretion kinetics of the corresponding GSH conjugates were studied in the rat in vivo after administration of (R)- or (S)-BPP at a dose level of 50 mumol/kg. Elimination of (R)-BPP from blood was faster than that of (S)-BPP: half lives were 9 +/- 2 min for (R)-BPP and 13 +/- 1 min for (S)-BPP. The biliary excretion rate of the GSH conjugate of (R)-BPP declined monoexponentially, while that of the GSH conjugate of (S)-BPP displayed a biphasic profile. Half lives of excretion were 13 +/- 1 for the GSH conjugate of (R)-BPP, and 11 +/- 2 for the GSH conjugate of (S)-BPP (second phase). The first phase in the biliary excretion of the GSH conjugate of (S)-BPP could not be attributed to capacity limitation of biliary transport carriers as higher excretion rates were attained upon administration of higher doses (100 and 200 mumol/kg) of (S)-BPP). The blood elimination profiles of (R)- and (S)-BPP differed greatly from the biliary excretion profiles of the corresponding GSH conjugates. This suggests that the kinetics of BPP conjugate excretion are determined by other processes than hepatic GSH conjugation.  相似文献   

17.
2-Methyl-branched chain acyl-CoA dehydrogenase was purified to homogeneity from rat liver mitochondria. The native molecular weight of the enzyme was estimated to be 170,000 by gel filtration. On sodium dodecyl sulfate-polyacrylamide gel electrophoresis both with and without 2-mercaptoethanol, the enzyme showed a single protein band with Mr = 41,500, suggesting that this enzyme is composed of four subunits of equal size. Its isoelectric point was 5.50 +/- 0.2, and A1%280 nm was 12.5. This enzyme contained protein-bound FAD. The purified enzyme dehydrogenated S-2-methylbutyryl-CoA and isobutyryl-CoA with equal activity. The activities with each of these compounds were co-purified throughout the entire purification procedure. This enzyme also dehydrogenated R-2-methylbutyryl-CoA, but the specific activity was considerably lower (22%) than that for the S-enantiomer. The enzyme did not dehydrogenate other acyl-CoAs, including isovaleryl-CoA, propionyl-CoA, butyryl-CoA, octanoyl-CoA, and palmitoyl-CoA, at any significant rate. Apparent Km and Vmax values for S-2-methylbutyryl-CoA were 20 microM and 2.2 mumol min-1 mg-1, respectively, while those for isobutyryl-CoA were 89 microM and 2.0 mumol min-1 mg-1 using phenazine methosulfate as an artificial electron acceptor. The enzyme was also active with electron transfer flavoprotein. Tiglyl-CoA and methacrylyl-CoA were identified as the reaction products from S-2-methylbutyryl-CoA and isobutyryl-CoA, respectively. 2-Ethylacrylyl-CoA was produced from R-2-methylbutyryl-CoA. Tiglyl-CoA competitively inhibited the activity with both S-2-methylbutyryl-CoA and isobutyryl-CoA with a similar Ki. The enzyme activity was also severely inhibited by several organic sulfhydryl reagents such as N-ethylmaleimide, p-hydroxymercuribenzoate, and methyl mercury iodide. The pattern and degree of inhibition were essentially identical for both substrates. The purified 2-methyl-branched chain acyl-CoA dehydrogenase was immunologically distinct from isovaleryl-CoA-, short chain acyl-CoA-, medium chain acyl-CoA-, or long chain acyl-CoA dehydrogenase.  相似文献   

18.
Rat hepatic microsomal squalene synthetase (EC 2.5.1.21) was induced 25-fold by feeding rats with diet containing the hydroxymethylglutaryl-coenzyme A reductase inhibitor, fluvastatin, and cholestyramine, a bile acid sequestrant. A soluble squalene synthetase protein with an estimated mass of 32-35 kDa, as determined by gel filtration chromatography on Sephacryl S-200 column, was solubilized out of the microsomes by controlled proteolysis with trypsin. Approximately 25% of the activity was recovered in a soluble form. The enzyme was purified to homogeneity utilizing a series of column chromatography purification steps on DEAE-cellulose, hydroxylapatite, and phenyl-Sepharose sequentially. The purified enzyme showed a single band on sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Initial kinetic analysis indicated an S0.5 values for trans-farnesyl diphosphate of 1.0 microM and for NADPH of 40 microM. The Vmax with respect to trans-farnesyl diphosphate was calculated at 1.2 mumol/min/mg. NADH also serves as substrate for the reaction with S0.5 value of 800 microM. Western blot analysis utilizing rabbit antisera raised against the purified, trypsin-truncated enzyme showed a single band for the isolated solubilized enzyme at 32-33 kDa and a band for the intact microsomal enzyme at about 45-47 kDa.  相似文献   

19.
P Le Maréchal  C Froussios  R Azerad 《Biochimie》1986,68(10-11):1211-1215
(3R) and (3S) 3-deoxy-3-fluoro-7-phospho-D-arabino hept-2-ulosonic acids (3R and 3S-3F-DAHP) the 3-fluoro analogues of DAHP were synthesized from the corresponding 2-deoxy-2-fluoro hexose-6-phosphates. 3R- and 3S-3F-DAHP were tested as substrates for 3-dehydroquinate synthetase from E. coli. Determination of kinetic parameters showed that their apparent Km and Vm were in the same order of magnitude for these two compounds. Further conversion of 3R- and 3S-3F-DAHP into (6R) and (6S) 6-fluoro dehydroshikimate and (6R) and (6S) 6-fluoro shikimate, respectively, was investigated and results are discussed.  相似文献   

20.
Two soluble forms of inositol phosphate 5-phosphomonoesterase have been partially purified and characterized from rat brain and are referred to as type 1 and type 2 according to their order of elution from DEAE-Sepharose. Together, these enzymes represent 26 +/- 3% (mean +/- S.E., n = 4) of the total inositol 1,4,5-triphosphate (Ins(1,4,5)P3) phosphatase activity assayed in crude brain homogenate and are present in approximately equal total activities in a 100,000 x g supernatant, with the remainder being membrane-bound. Both soluble enzymes require Mg2+ for activity, are moderately inhibited by Ca2+ in the micromolar range, and can be inhibited by millimolar concentrations of a variety of phosphorylated compounds. The type 1 enzyme has been purified to a specific activity of 1.06 mumol/min/mg protein. It elutes as a 60-kDa protein on Sephacryl S-200. On sodium dodecyl sulfate-polyacrylamide gel electrophoresis, the type 1 enzyme correlates with a pair of protein bands of 66 and 60 kDa. It has apparent Km values of 3 and 0.8 microM for Ins(1,4,5)P3 and inositol 1,3,4,5-tetrakisphosphate (Ins(1,3,4,5)P4), respectively, and hydrolyses Ins(1,4,5)P3 approximately 12 times faster than Ins(1,3,4,5)P4. The type 2 enzyme has been purified to a specific activity of 15.2 mumol/min/mg protein, elutes as a protein of 160 kDa on Sephacryl S-300, and migrates as a similarly sized subunit on sodium dodecyl sulfate-polyacrylamide gel electrophoresis. It has an apparent Km for Ins(1,4,5)P3 of 18 microM. Its apparent Km for Ins(1,3,4,5)P4, however, is greater than 150 microM, suggesting that this enzyme is primarily an Ins(1,4,5)P3 5-phosphomonoesterase. The relationship of these two enzymes to the inositol tris/tetrakisphosphate pathway is discussed.  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