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 effect of temperature on enantioselectivity and desymmetrization in the acetylation of 2-phenyl-1,3-propanediol (1a), 2-benzyl-1,3-propanediol (1b), 2-methyl-2-phenyl-1,3-propanediol (1c) and 2-benzyl-2-methyl-1,3-propanediol (1d) by a novel lipase (CSL) isolated from the yeast Cryptococcus spp. S-2 was studied. Desymmetrization of 1a, 1c and 1d by CSL-catalyzed acetylation was observed in the temperature range of -20°C to 40°C, while diacetylation of 1b occurred considerably even at 0°C. The kinetic parameters of the selectivity indicated that the acetylation of 1a is an entropy controlled process whereas the reaction of 1c and 1d is mainly controlled by the enthalpy term. In the monoacetylation of the diol 1d, the preferred configuration in the enantiomeric induction by CSL was opposite to that of immobilized porcine pancreatic lipase (PPL).  相似文献   

2.
The kinetics of the biodegradation of two commercial plasticizers, diethylene glycol dibenzoate (D(EG)DB) and dipropylene glycol dibenzoate (D(PG)DB), as well as two alternative plasticizers, 1,3-propanediol dibenzoate and 2,2-methyl-propyl-1,3-propanediol dibenzoate, were investigated in an aerated bioreactor. The experiments were conducted with resting cells of Rhodococcus rhodochrous, which had been grown with hexadecane as the substrate. The first step in the biodegradation was always the hydrolysis of an ester bond, releasing the corresponding monobenzoate and benzoic acid. Biodegradation of plasticizers and their associated metabolites were modeled using a Monod-type kinetic model. Significant differences between the biodegradation of commercial and alternative plasticizers were observed both in the biodegradation pathway and the biodegradation rates of monobenzoate metabolites. At a selected concentration of 0.4 g/L, the monobenzoates released from the biodegradation of 1,3-propanediol dibenzoate and 2,2-methyl-propyl-1,3-propanediol dibenzoate were degraded approximately 13 and 4 times more quickly, respectively, than the monobenzoate released from the biodegradation of D(PG)DB. The rapid biodegradation of monobenzoates released from microbial hydrolysis of alternative dibenzoate plasticizers was attributed to the lack of an ether bond in these compounds.  相似文献   

3.
Abstract In a mineral medium containing sulfate as terminal electron acceptor, the sulfate-reducing bacterium Desulfovibrio alcoholovorans oxidized stoichiometrically 1 mol glycerol to 1 mol acetate and 1 mol 1,3-propanediol to 1 mol acetate with the concomitant reduction of 0.75 and 1 mol sulfate, respectively; 1 mol 1,2-propanediol was degraded to 0.8 mol acetate and 0.1 mol proprionate, with the reduction of approximately 1 mol sulfate. The maximum specific growth rates (μmax in h−1) were 0.22, 0.086 and 0.09 with glycerol, 1,3-propanediol and 1,2-propanediol, respectively. The growth yields were 12.7 g, 11.1 g and 7.2 g dry weight/mol 1,3-propanediol, glycerol and 1,2-propanediol degraded, respectively. The growth yields and maximum specific growth rates of the H2-transferring associations were also calculated. In the absense of sulfate, all these reduced substrates were degraded to acids and methane when D. alcoholovorans was cocultured with Methanospirillum hungatei . Changes in the metabolic pathway were observed in the degradation of 1,2- and 1,3-propanediol. The metabolic efficiency of D. alcoholovorans to degrade glycerol, 1.2- and 1,3-propanediol is discussed.  相似文献   

4.
The absolute configuration at C-12 of pittosporatobiraside A and B isolated from the leaves of Pittosporum tobira was determined to be S on the basis of the exciton chirality of their dibenzoate derivative. The structures of the two glycosides were thus established to be (1S,9S,10S,11S,12S,14R,16R)-12-[(Z)-2-methyl-1-oxo-2-butenyl]-6,14-dimethyl-2-methylene-9-(1-methylethyl)-15,17-dioxatricyclo[8.7.0.011,16]heptadec-5-en-13-one and (1S,9S,10S,11S,12S,14R,16R)-12-(3-methyl-1-oxo-2-butenyl)-6,14-dimethyl-2-methylene-9-(1-methylethyl)-15,17-dioxatricyclo [8.7.0.011,16]heptadec-5-en-13-one, respectively.  相似文献   

5.
The 1,3-regiospecific lipase from Candida deformanscatalysed the esterification of oleic acid and propanediol in biphasic aqueous/lipid medium without organic solvent. The highest conversion of oleic acid into 1,2-propanediol ester was 74% in 24 h with 6.25 mol/l 1,2-propanediol and 0.08 mol/l oleic acid, and produced 100% monoester. The esterification of 1,3-propanediol converted up to 98% of oleic acid into esters in 24 h (with 7.5 mol/l 1,3-propanediol and 0.08 mol/l oleic acid) and formed 35-90% monoester depending on 1,3-propanediol initial concentration (2.5-10 mol/l).  相似文献   

6.
Reactive extraction for downstream separation of 1,3-propanediol   总被引:4,自引:0,他引:4  
The downstream separation of 1,3-propanediol from dilute aqueous solution was studied. A process combining reversible reaction of 1, 3-propanediol with acetaldehyde to 2-methyl-1,3-dioxane and a simultaneous extraction of the product by organic solvent appears to be technically feasible and attractive. The dioxane yield was 91-92%, the overall conversion of 1,3-propanediol was ca. 98%, and recovery of dioxane into the organic extractant was 75%.  相似文献   

7.
The new PN ligands 5, 6 and 7 were prepared by Schiff base condensation of 2-formylphenyl(diphenyl)phosphine (1) with the optically active amines (R)-(−)-2-aminobutanol (2), (S)-(+)-2-aminobutanol (3) and (1S,2S)-2-amino- 1-phenyl-1,3-propanediol (4). These new ligands were used in the Pd catalysed allylation of 1,5-dimethylbarbituric acid with allylacetate. 5-Allyl-1,5-dimethylbarbituric acid was obtained with an optical induction of up to 12.7% ee.  相似文献   

8.
Cholic and deoxycholic acid amides 10-17 have been synthesised from (1R,2R)-1-phenyl-2-amino-1,3-propanediol 2, (1S,2S)-1-phenyl-2-amino-1,3-propanediol 4, (1R,2R)-1-para-nitrophenyl-2-amino-1,3-propanediol 3, (1S,2S)-1-para-nitrophenyl-2-amino-1,3-propanediol 5. Amide 12 derived from N-succinimidyl ester 9 of deoxycholic acid and (1R,2R)-1-phenyl-2-amino-1,3-propanediol 2, found to be active against Cryptococcus neoformans and the amide 17 obtained from N-succinimidyl ester 9 of deoxycholic acid and (1S,2S)-1-para-nitrophenyl-2-amino-1,3-propanediol 5, is found to be potent against various gram-positive bacteria.  相似文献   

9.
P Boutron 《Cryobiology》1992,29(3):347-358
A 2,3-butanediol containing 96.7% (w/w) racemic mixture of the levo and dextro isomers and only 3.1% (w/w) of the meso isomer (called 2,3-butanediol 97% dl) has been used for the cryoprotection of red blood cells. The erythrocytes were cooled to -196 degrees C at rates between 2 and 3500 degrees C/min, followed by slow or rapid warming. Up to 20% (w/w) of this polyalcohol, only the classical peak of survival is observed, as with up to 20% (w/w) 1,2-propanediol or 1,3-butanediol. Twenty percent 2,3-butanediol 97% dl can protect red blood cells very efficiently. The maximum survival, of 90%, as with 20% glycerol, is a little lower than with 20% 1,2-propanediol and higher than with 20% 1,3-butanediol. Fifteen percent 2,3-butanediol protects fewer red blood cells than 15% glycerol or 1,2-propanediol, with a maximum survival of about 80%. The best cryoprotection by 30% 2,3-butanediol 97% dl is obtained at the slowest cooling and warming rates, where survival approaches 90%. After a minimum, an increase of survival is observed at the fastest cooling rates, which would correspond to complete vitrification. These rates are lower than with 30%, 1,2-propanediol or 1,3-butanediol, in agreement with the higher glass-forming tendency of 2,3-butanediol 97% dl solutions. In agreement with the remarkable physical properties of its aqueous solutions, the present experiments also suggest that 2,3-butanediol containing mainly the levo and dextro isomers could be a very useful cryoprotectant for organ cryopreservation. However, it would perhaps be better to use it in combination with other cryoprotectants, since it is a little more toxic than glycerol or 1,2-propanediol at high concentrations.  相似文献   

10.
Glycerol can be oxidized to formaldehyde by rat liver microsomes and by cytochrome P450. The ability of other alcohols to be oxidized to formaldehyde was determined to evaluate the structural determinants of the alcohol which eventually lead to this production of formaldehyde. Monohydroxylated alcohols such as 1- or 2-propanol did not produce formaldehyde when incubated with NADPH and microsomes. Geminal diols such as 1,3-propanediol, 1,3-butanediol, or 1,4-butanediol also did not yield formaldehyde. However, vicinal diols such as 1,2-propanediol or 1,2-butanediol produced formaldehyde. With 1,2-propanediol, the residual two-carbon fragment was found to be acetaldehyde, while with 1,2-butanediol, the residual three-carbon fragment was propionaldehyde. Oxidation of 1,2-propanediol to formaldehyde plus acetaldehyde involved interaction with an oxidant derived from H2O2 plus nonheme iron, since production of the two aldehydic products was completely prevented by catalase or glutathione plus glutathione peroxidase and by chelators such as desferrioxamine or EDTA. The oxidant was not superoxide or hydroxyl radical. Product formation was fivefold lower when NADH replaced NADPH, and was inhibited by substrates, ligands, and inhibitors of cytochrome P450. A charged glycol such as alpha-glycerophosphate (but not the geminal beta-glycerophosphate) was readily oxidized to formaldehyde, suggesting that interaction of the glycol with the oxidant was occurring in solution and not in a hydrophobic environment. These results indicate that the carbon-carbon bond between 1,2-glycols can be cleaved by an oxidant derived from microsomal generated H2O2 and reduction of non-heme iron, with the subsequent production of formaldehyde plus an aldehyde with one less carbon than the initial glycol substrate.  相似文献   

11.
The effect of temperature on enantioselectivity and desymmetrization in the acetylation of 2-phenyl-1,3-propanediol (1a), 2-benzyl-1,3-propanediol (1b), 2-methyl-2-phenyl-1,3-propanediol (1c) and 2-benzyl-2-methyl-1,3-propanediol (1d) by a novel lipase (CSL) isolated from the yeast Cryptococcus spp. S-2 was studied. Desymmetrization of 1a, 1c and 1d by CSL-catalyzed acetylation was observed in the temperature range of ?20°C to 40°C, while diacetylation of 1b occurred considerably even at 0°C. The kinetic parameters of the selectivity indicated that the acetylation of 1a is an entropy controlled process whereas the reaction of 1c and 1d is mainly controlled by the enthalpy term. In the monoacetylation of the diol 1d, the preferred configuration in the enantiomeric induction by CSL was opposite to that of immobilized porcine pancreatic lipase (PPL).  相似文献   

12.
All the aqueous solutions of linear saturated polyalcohols with four carbons have been investigated at low temperature. Only ice has been observed in the solutions of 1,3-butanediol and 1,2,3- and 1,2,4-butanetriol. For same solute concentration, the glass-forming tendency on cooling is highest with 2,3-butanediol, where it is comparable to that with 1,2-propanediol, the best solute reported to date. However, the quantity of ice and hydrate crystallized is particularly high on slow cooling or on subsequent rewarming. The highest stability of the amorphous state is observed on rewarming the 1,2-butanediol and 1,3-butanediol solutions. With respect to this property, these compounds come just after 1,2-propanediol and before all the other compounds studied so far. They are followed by dimethylsulfoxide and 1,2,3-butanetriol. The glass-forming tendency of the 1,3-butanediol solutions is also very high; it is third only to that of 1,2-propanediol and 2,3-butanediol. The glass-forming tendency is a little smaller with 1,2-butanediol, but it is cubic instead of ordinary hexagonal ice which crystallizes on cooling rapidly with 35% 1,2-butanediol. Cubic ice is thought to be innocuous. A gigantic glass transition is observed with 45% of this strange solute. 1,4-Butanediol, 45% also favors cubic ice greatly. Therefore, 1,2- and 1,3-butanediol with comparable physical properties are perhaps as interesting as 1,2-propanediol for cryopreservation of cells or organs by complete vitrification. Together with 1,2-propanediol, 1,2- and 1,3-butanetriol, 1,2,3-butanetriol, and perhaps 2,3-butanediol provide an interesting battery of solutions for cryopreservation by vitrification.  相似文献   

13.
Rhodococcus sp. 2N was found as a 1,3-propanediols-oxidizing strain from soil samples through enrichment culture using 2,2-diethyl-1,3-propanediol (DEPD) as the sole carbon source. The culture condition of the strain 2N was optimized, and the highest activity was observed when 0.3% (w/v) DEPD was added in the culture medium as an inducer. Chiral HPLC analysis of the hydroxyalkanoic acid converted from 2-ethyl-2-methyl-1,3-propanediol (EMPD) revealed that the strain 2N catalyzed the (R)-selective oxidation of EMPD. The reaction products and intermediates from DEPD and EMPD were identified by nuclear magnetic resonance analyses, and the results suggested that only one hydroxymethyl group of the propanediols was converted to carboxy group via two oxidation steps. Under optimized conditions and after a 72-h reaction time, the strain 2N produced 28 mM (4.1 g/L) of 2-(hydroxymethyl)-2-methylbutanoic acid from EMPD with a molar conversion yield of 47% and 65% ee (R).  相似文献   

14.
Whole cells of Rhodococcus erythropolis DSM 44534 grown on ethanol, (R)- and (S)-1,2-propanediol were used for biotransformation of racemic 1,4-alkanediols into γ-lactones. The cells oxidized 1,4-decanediol (1a) and 1,4-nonanediol (2a) into the corresponding γ-lactones 5-hexyl-dihydro-2(3H)-furanone (γ-decalactone, 1c) and 5-pentyl-dihydro-2(3H)-furanone (γ-nonalactone, 2c), respectively, with an EE(R) of 40–75%. The transient formation of the γ-lactols 5-hexyl-tetrahydro-2-furanol (γ-decalactol, 1b) and 5-pentyl-tetrahydro-2-furanol (γ-nonalactol, 2b) as intermediates was observed by GC–MS. 1,4-Pentanediol (3a) was transformed into 5-methyl-dihydro-2(3H)-furanone (γ-valerolactone, 3c) whereas (R)- and (S)-2-methyl-1,4-butanediol (4a) was converted to the methyl-substituted γ-butyrolactones 4-methyl-dihydro-2(3H)-furanone (4c1) and 3-methyl-dihydro-2(3H)-furanone (4c2) in a ratio of 80:20 with a yield of 55%. Also cis-2-buten-1,4-diol (5a) was transformed resulting in the formation of 2(5H)-furanone (γ-crotonolactone, 5c). At the higher pH values of 8.8 the yield of lactone formed was improved; however, the enatiomeric excesses were slightly higher at the lower pH of 5.2.  相似文献   

15.
The haloalkane dehalogenases are detoxifying enzymes that convert a broad range of halogenated substrates to the corresponding alcohols. Complete crystal structures of haloalkane dehalogenase from Sphingomonas paucimobilis UT26 (LinB), and complexes of LinB with 1,2-propanediol/1-bromopropane-2-ol and 2-bromo-2-propene-1-ol, products of debromination of 1,2-dibromopropane and 2,3-dibromopropene, respectively, were determined from 1.8 A resolution X-ray diffraction data. Published structures of native LinB and its complex with 1,3-propanediol [Marek et al. (2000) Biochemistry 39, 14082-14086] were reexamined. The full and partial debromination of 1,2-dibromopropane and 2,3-dibromopropene, respectively, conformed to the observed general trend that the sp(3)-hybridized carbon is the predominant electrophilic site for the S(N)2 bimolecular nucleophilic substitution in dehalogenation reaction. The 2-bromo-2-propene-1-ol product of 2,3-dibromopropene dehalogenation in crystal was positively identified by the gas chromatography-mass spectroscopy (GC-MS) technique. The 1,2-propanediol and 1-bromopropane-2-ol products of 1,2-dibromopropane dehalogenation in crystal were also supported by the GC-MS identification. Comparison of native LinB with its complexes showed high flexibility of residues 136-157, in particular, Asp146 and Glu147, from the cap domain helices alpha(4) and alpha(5)('). Those residues were shifted mainly in direction toward the ligand molecules in the complex structures. It seems the cap domain moves nearer to the core squeezing substrate into the active center closer to the catalytic triad. This also leads to slight contraction of the whole complex structures. The flexibility detected by crystallographic analysis is in remarkable agreement with flexibility observed by molecular dynamic simulations.  相似文献   

16.
An enzyme capable of dehalogenating vicinal haloalcohols to their corresponding epoxides was purified from the 3-chloro-1,2-propanediol-utilizing bacterium Arthrobacter sp. strain AD2. The inducible haloalcohol dehalogenase converted 1,3-dichloro-2-propanol, 3-chloro-1,2-propanediol, 1-chloro-2-propanol, and their brominated analogs, 2-bromoethanol, as well as chloroacetone and 1,3-dichloroacetone. The enzyme possessed no activity for epichlorohydrin (3-chloro-1,2-epoxypropane) or 2,3-dichloro-1-propanol. The dehalogenase had a broad pH optimum at about 8.5 and a temperature optimum of 50 degrees C. The enzyme followed Michaelis-Menten kinetics, and the Km values for 1,3-dichloro-2-propanol and 3-chloro-1,2-propanediol were 8.5 and 48 mM, respectively. Chloroacetic acid was a competitive inhibitor, with a Ki of 0.50 mM. A subunit molecular mass of 29 kDa was determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. With gel filtration, a molecular mass of 69 kDa was found, indicating that the native protein is a dimer. The amino acid composition and N-terminal amino acid sequence are given.  相似文献   

17.
Escherichia coli were grown on 14.3% uniformly 13C-labeled glucose as the sole carbon source and challenged anaerobically with 90% 13C-labeled formaldehyde. The major multiply labeled metabolites were identified by 13C NMR spectroscopy to be glycerol and 1,2-propanediol, and a minor metabolite was shown to be 1,3-propanediol. In each case, formaldehyde is incorporated only into the C1 position. A novel form of 13C NMR isotope dilution analysis of the major products reveals that all the 1,2-diol C1 is formaldehyde derived but that about 40% of the glycerol C1 is derived from bacterial sources. Glycerokinase converted the metabolite [1-13C]glycerol to equal amounts of [3-13C]glycerol 3-phosphate and [1-13C]glycerol 3-phosphate, demonstrating that the metabolite is racemic. When [13C]formaldehyde incubation was carried out in H2O/D2O mixtures, deuterium incorporation was detected by beta- and gamma-isotope shifts. The 1,3-diol is deuterium labeled only at C2 and only once, while the 1,2-diol and glycerol are each labeled independently at both C2 and C3; C3 is multiply labeled. Deuterium incorporation levels are different for each metabolite, indicating that the biosynthetic pathways probably diverge early.  相似文献   

18.
Fermentation of glycerol to 1,3-propanediol: use of cosubstrates   总被引:16,自引:0,他引:16  
Three fermentable substances, glucose, 1,2-ethanediol and 1,2-propanediol were checked as cosubstrates for the fermentation of glycerol by Clostridium butyricum and Citrobacter freundii with the aim of achieving a complete conversion of glycerol to 1,3-propanediol. Glucose was fermented by C. butyricum mainly to acetate, CO2 and reducing equivalents in the presence of glycerol and contributed markedly to the 1,3-propanediol yield. However, because of relatively slow growth on glucose, complete conversion was not achieved. If the two glycols were used as cosubstrates for glycerol fermentation, the 1,3-propanediol yield did not increase but dimished considerably, as they were converted to more reduced products, i.e. alcohols instead of acids. From 1,2-propanediol 2-propanol was formed in addition to 1-propanol. The ratio of the propanols was dependent on the culture conditions.  相似文献   

19.
A nicotinamide adenine dinucleotide-linked dehydrogenase has been partially purified from a mutant of Escherichia coli K-12 able to grow on l-1,2-propanediol as carbon and energy source. This enzyme catalyzes the dehydrogenation at carbon 1 of l-1,2-propanediol, glycerol, 1,3-propanediol, ethylene glycol, and ethyl alcohol. The purified protein requires added ferrous or managanous ions. The V(max) and the apparent K(m) for a given substrate vary with the particular metal used.  相似文献   

20.
An aerobic bacterium (Mycobacterium sp. strain ELW1) that utilizes 2-methylpropene (isobutylene) as a sole source of carbon and energy was isolated and characterized. Strain ELW1 grew on 2-methylpropene (growth rate = 0.05 h−1) with a yield of 0.38 mg (dry weight) mg 2-methylpropene−1. Strain ELW1 also grew more slowly on both cis- and trans-2-butene but did not grow on any other C2 to C5 straight-chain, branched, or chlorinated alkenes tested. Resting 2-methylpropene-grown cells consumed ethene, propene, and 1-butene without a lag phase. Epoxyethane accumulated as the only detected product of ethene oxidation. Both alkene consumption and epoxyethane production were fully inhibited in cells exposed to 1-octyne, suggesting that alkene oxidation is initiated by an alkyne-sensitive, epoxide-generating monooxygenase. Kinetic analyses indicated that 1,2-epoxy-2-methylpropane is rapidly consumed during 2-methylpropene degradation, while 2-methyl-2-propen-1-ol is not a significant metabolite of 2-methylpropene catabolism. Degradation of 1,2-epoxy-2-methylpropane by 2-methylpropene-grown cells led to the accumulation and further degradation of 2-methyl-1,2-propanediol and 2-hydroxyisobutyrate, two sequential metabolites previously identified in the aerobic microbial metabolism of methyl tert-butyl ether (MTBE) and tert-butyl alcohol (TBA). Growth of strain ELW1 on 2-methylpropene, 1,2-epoxy-2-methylpropane, 2-methyl-1,2-propanediol, and 2-hydroxyisobutyrate was fully inhibited when cobalt ions were omitted from the growth medium, while growth on 3-hydroxybutyrate and other substrates was unaffected by the absence of added cobalt ions. Our results suggest that, like aerobic MTBE- and TBA-metabolizing bacteria, strain ELW1 utilizes a cobalt/cobalamin-dependent mutase to transform 2-hydroxyisobutyrate. Our results have been interpreted in terms of their impact on our understanding of the microbial metabolism of alkenes and ether oxygenates.  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