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.
Asymmetric hydrolytic enzymes for trans-3-(4-methoxyphenyl)glycidic acid methyl ester [(±)-MPGM], a key intermediate in the synthesis of diltiazem hydrochloride that is useful as a coronary vasodilator, were screened from 730 microorganisms. Among the microbial enzymes tested, Serratia marcescens lipase had the highest enantioselectivity (E=135) for hydrolysis of (±)-MPGM in a two-phase system using water (pH 8) and a water-immiscible solvent such as toluene. Resolution of (±)-MPGM by S. marcescens lipase gave (2R, 3S)-3-(4-methoxyphenyl)glycidic acid methyl ester [(−)-MPGM] with a reaction yield of 48% and optical purity of >99.9% e.e. After the reaction, the emulsion of toluene and water was separated into two clear layers by the addition of sodium dodecyl sulfate. The crystalline (−)-MPGM was isolated with a yield of over 43% and optical purity of 100% e.e. without column treatment. Diltiazem hydrochloride synthesis using asymmetric hydrolysis of (±)-MPGM was found to be a more efficient process compared to the conventional chemical synthetic process. Some enzymatic characterizations on asymmetric hydrolysis in two-phases of organic solvent-water by S. marcescens lipase were investigated.  相似文献   

2.
(R,S)-Isoproterenol inhibits the formation of toxic granular tau oligomers associated with neuronal loss and development of cognitive disorders, and is an attractive drug candidate for Alzheimer’s disease. To elucidate its behavior in the brain by positron emission tomography, we synthesize (R,S)-[11C]isoproterenol by reductive alkylation of (R,S)-norepinephrine with [2-11C]acetone, which was in turn synthesized in situ under improved conditions afforded a decay-corrected radiochemical yield of 54%. The reductive alkylation using NaBH(OAc)3 as reducing agent in the presence of benzoic acid in DMSO/DMF (60:40 v/v) at 100 °C for 10 min gave (R,S)-[11C]isoproterenol in an 87% radio-high performance liquid chromatography (HPLC) analytical yield. HPLC separation using a strong cation exchange column, followed by pharmaceutical formulation in the presence of d/l-tartaric acid, afforded (R,S)-[11C]isoproterenol with a total radioactivity of 2.0 ± 0.2 GBq, a decay-corrected radiochemical yield of 19 ± 2%, chemical and radiochemical purities of 71% and >99%, respectively, and a molar activity of 100 ± 13 GBq/μmol (n = 3). The overall synthesis time from the end of the bombardment to pharmaceutical formulation was 48 min. A preliminary preclinical PET study in a rat demonstrated the potential of the radioligand for the evaluation of the penetration of (R,S)-isoproterenol in human brain.  相似文献   

3.
Proteolytic and cataplerotic sources of hepatic glutamine were determined by 2H NMR analysis of urinary phenylacetylglutamine (PAGN) 2H-enrichments in eight healthy subjects after 2H2O and phenylbutyric acid ingestion. Body water enrichment was 0.49±0.03%. PAGN was enriched to lower levels with significant differences between the various glutamine positions. PAGN position 2 enrichment=0.33±0.02%; 3R=0.27±0.02%; 3S=0.27±0.02% and position 4=0.17±0.01%. Position 3R,S enrichments are conditional with the net conversion of citrate to glutamate and are therefore markers of cataplerosis. From the ratio of positions 3R,S to body water enrichment, 55±3% of hepatic glutamine was derived from cataplerosis and 45±3% from proteolysis. In conclusion, enrichment of PAGN 3R,S hydrogens relative to that of body water reflects the contribution of cataplerotic and proteolytic sources to hepatic glutamine.  相似文献   

4.
The lipase from Serratia marcescence ECU1010 (Sml) was capable of enantioselectively catalyzing the synthesis of many chiral drug precursors. This paper investigated the immobilization of Sml on appropriate supporting materials and its performance in bioreactor. Chitosan, Celite 545, and DEAE-cellulose were found to be the ideal supports among 8 carriers tested with respect to enzyme load and activity recovery of lipase. When Sml was immobilized, significant improvements of stability against pH, thermal, and operational deactivation were observed with all the 3 better supports, and the best stability was observed when the lipase was immobilized on glutaraldehyde activated chitosan. As for the effect of organic solvent in the biphasic reaction system, the hydrolytic activity of the immobilized lipase on trans-3-(4′-methoxyphenyl)glycidic acid methyl ester ((±)-MPGM) observed in isopropyl ether was 6 and 3 times higher than those in toluene and methyl tert-butyl ether. The lipasecatalyzed production of (−)-MPGM by enzymatic resolution of (±)-MPGM with chitosan-Sml in isopropyl etherwater biphasic system was carried out in a 2 L stirred-tank reactor. The batch operation was more efficient operation mode for the enantioselective hydrolysis of (±)-MPGM, affording enantiopure (−)-MPGM in 44.3% overall yield, in contrast to 29.3% in a continuous reactor.  相似文献   

5.
In the Candida antarctica lipase B-catalyzed hydrolysis of (R,S)-azolides derived from (R,S)-N-protected proline in water-saturated methyl tert-butyl ether (MTBE), high enzyme activity with excellent enantioselectivity (V S V R ?1 ?>?100) for (R,S)-N-Cbz-proline 1,2,4-triazolide (1) and (R,S)-N-Cbz-proline 4-bromopyrazolide (2) was exploited in comparison with their corresponding methyl ester analog (3). Changing of the substrate structure, water content, solvent, and temperature was found to have profound influences on the lipase performance. On the basis of enzyme activity and enantioselectivity and solvent boiling point, the best reaction condition of using 1 as the substrate in water-saturated MTBE at 45 °C was selected and further employed for the successful resolution of (R,S)-N-Cbz-pipecolic 1,2,4-triazolide (5) and (R,S)-N-Boc-nipecotic 1,2,4-triazolide (9). Moreover, more than 89.1 % recovery of remained (R)-1 is obtainable in five cycles of enzyme reusage, when pH 7 phosphate buffers were employed as the extract at 4 °C.  相似文献   

6.
New neprilysin inhibitors containing an α-mercaptoketone HSC(R1R2)CO group, as zinc ligand were designed. Two parameters were explored for potency optimization: the size of the inhibitor which could interact with the S1, S1′ or S2′ domain of the enzyme and the nature of the substituents R1, R2 of the mercaptoketone group. Introduction of a cyclohexyl chain in R1, R2 position and a (3-thiophen)benzyl group in position R3 (compound 12n) yielded to the most potent inhibitor of this series with a Ki value of 2 ± 0.3 nM. This result suggests that this new inhibitor interacts within the S1, S1′ domain of NEP allowing a pentacoordination of the catalytic Zn2+ ion by the mercaptoketone moiety.  相似文献   

7.
This study examined the role and physiological relevance of 3-hydroxyisobutyrate dehydrogenase-I (3HIBDHI) of Pseudomonas denitrificans ATCC 13867 in the degradation of 3-hydroxypropionic acid (3-HP) during 3-HP production. The gene encoding 3HIBDH-I of P. denitrificans ATCC 13867 was cloned and expressed in Escherichia coli BL21 (DE3). The recombinant 3HIBDH-I was then purified on a Ni-NTA-HP column and characterized for its choice of substrates, cofactors, metals, reductants, and the optimal temperature and pH. The recombinant 3HIBDH-I showed a high catalytic constant (k cat/K m) of 604.1 ± 71.1 mM/S on (S)-3-hydroxyisobutyrate, but no detectable activity on (R)-3-hydroxyisobutyrate. 3HIBDH-I preferred NAD+ over NADP+ as a cofactor for its catalytic activity. The k cat/K m determined for 3-HP was 15.40 ± 1.43 mM/S in the presence of NAD+ at 37°C and pH 9.0. In addition to (S)-3-hydroxyisobutyrate and 3-HP, 3HIBDH-I utilized l-serine, methyl-d,l-serine, and methyl-(S)-(+)-3-hydroxy-2-methylpropionate; on the other hand, the k cat/K m values determined for these substrates were less than 5.0mM/S. Ethylenediaminetetraacetic acid, 2-mercaptoethanol, dithiothreitol and Mn2+ increased the activity of 3HIBDHI significantly, whereas the presence of Fe2+, Hg2+ and Ag+ in the reaction mixture at 1.0 mM completely inhibited its activity. This study revealed the characteristics of 3HIBDH-I and its significance in 3-HP degradation.  相似文献   

8.
The meso-2,3-butanediol dehydrogenase (meso-BDH) from S. marcescens H30 is responsible for converting acetoin into 2,3-butanediol during sugar fermentation. Inactivation of the meso-BDH encoded by budC gene does not completely abolish 2,3-butanediol production, which suggests that another similar enzyme involved in 2,3-butanediol formation exists in S. marcescens H30. In the present study, a glycerol dehydrogenase (GDH) encoded by gldA gene from S. marcescens H30 was expressed in Escherichia coli BL21(DE3), purified and characterized for its properties. In vitro conversion indicated that the purified GDH could catalyze the interconversion of (3S)-acetoin/meso-2,3-butanediol and (3R)-acetoin/(2R,3R)-2,3-butanediol. (2S,3S)-2,3-Butanediol was not a substrate for the GDH at all. Kinetic parameters of the GDH enzyme showed lower K m value and higher catalytic efficiency for (3S/3R)-acetoin in comparison to those for (2R,3R)-2,3-butanediol and meso-2,3-butanediol, implying its physiological role in favor of 2,3-butanediol formation. Maximum activity for reduction of (3S/3R)-acetoin and oxidations of meso-2,3-butanediol and glycerol was observed at pH 8.0, while it was pH 7.0 for diacetyl reduction. The enzyme exhibited relative high thermotolerance with optimum temperature of 60 °C in the oxidation–reduction reactions. Over 60 % of maximum activity was retained at 70 °C. Additionally, the GDH activity was significantly enhanced for meso-2,3-BD oxidation in the presence of Fe2+ and for (3S/3R)-acetoin reduction in the presence of Mn2+, while several cations inhibited its activity, particularly Fe2+ and Fe3+ for (3S/3R)-acetoin reduction. The properties provided potential application for single configuration production of acetoin and 2,3-butanediol .  相似文献   

9.
The synthesis of methyl (?)-shikimate [(?)-2] was achieved via lipase-catalyzed optical resolution of (1S*,4R*,5R*)-4-hydroxy-6-oxabicyclo[3.2.1]oct-2-en-7-one (3). Transesterification of (±)-3 and vinyl acetate with lipase MY and subsequent hydrolysis gave optically pure (?)-3. This compound was converted to (?)-2 in two steps.  相似文献   

10.
Secoisolariciresinol diglucosides (SDGs) (S,S)-SDG-1 (major isomer in flaxseed) and (R,R)-SDG-2 (minor isomer in flaxseed) were synthesized from vanillin via secoisolariciresinol (6) and glucosyl donor 7 through a concise route that involved chromatographic separation of diastereomeric diglucoside derivatives (S,S)-8 and (R,R)-9. Synthetic (S,S)-SDG-1 and (R,R)-SDG-2 exhibited potent antioxidant properties (EC50 = 292.17 ± 27.71 μM and 331.94 ± 21.21 μM, respectively), which compared well with that of natural (S,S)-SDG-1 (EC50 = 275.24 ± 13.15 μM). These values are significantly lower than those of ascorbic acid (EC50 = 1129.32 ± 88.79 μM) and α-tocopherol (EC50 = 944.62 ± 148.00 μM). Compounds (S,S)-SDG-1 and (R,R)-SDG-2 also demonstrated powerful scavenging activities against hydroxyl [natural (S,S)-SDG-1: 3.68 ± 0.27; synthetic (S,S)-SDG-1: 2.09 ± 0.16; synthetic (R,R)-SDG-2: 1.96 ± 0.27], peroxyl [natural (S,S)-SDG-1: 2.55 ± 0.11; synthetic (S,S)-SDG-1: 2.20 ± 0.10; synthetic (R,R)-SDG-2: 3.03 ± 0.04] and DPPH [natural (S,S)-SDG-1: EC50 = 83.94 ± 2.80 μM; synthetic (S,S)-SDG-1: EC50 = 157.54 ± 21.30 μM; synthetic (R,R)-SDG-2: EC50 = 123.63 ± 8.67 μM] radicals. These results confirm previous studies with naturally occurring (S,S)-SDG-1 and establish both (S,S)-SDG-1 and (R,R)-SDG-2 as potent antioxidants and free radical scavengers for potential in vivo use.  相似文献   

11.
(1) Only (R,S)2′,3′-epoxypropyl β-d-glucopyranoside of the complete series of mono (R,S)2′.3′-epoxypropyl ethers and glycosides of d-glucopyranose significantly inactivated yeast hexokinase.(2) (R,S)2′,3′-Epoxypropyl β-d-glucopyranoside inactivates yeast hexokinase in the absence of MgATP2?, The rate of inactivation is unaffected by MgATP2?.(3) The rate of inactivation of hexokinase with (R,S)2′,3′-epoxypropyl β-d-ilucopyranoside was much greater when hexokinase was present in a monomeric form than when it was present in a dimeric form.(4) (R,S)2′,3′-Epoxypropyl β-d-glucopyranoside has a high Kt (0.38 M) and at a saturating concentrarion, the first order rate constant for the inactivation of monomeric hexokinase is 8.3 · 10?4 sec.(5) d-Glucose protects against this inactivation and this was used to derive a dissocistion constant of 0.21 mM for d-glucose in the absence of MgATP2?.(6) The alkylation of yeast hexokinase by (R,S)2′,3′-epoxypropyl β-d-gluco-pyranoside was not specific to the active site. When the concentration of (R,S)2′,3′-epoxypropyl β-d-glucopyranoside was 50 mM two thiol groups outside the active site were also alkylated.(7) The reaction between 5,5′-dithiobis-(2-nitrobenzoic acid) (DTNB) and yeast hexokinase was examined in detail. Two thiol groups per monomer (mol. wt. 50000) reacted with a second order rate constant of 27 1 mole?1 sec?1. A third thiol group reacted more slowly with a second-order rate constant of 1.6 1 mole?1 sec?1 and a fourth thiol group reacted very slowly with inactivation of the enzyme. Tue second-order rate constant in this case was 0.1 1 mole?1 sec?1.  相似文献   

12.
Production and activity of extracellular lipase from Luteibacter sp.   总被引:1,自引:0,他引:1  
Microbial lipases are widely used in industrial applications due to their versatility, and the characterization of new lipase-producing microorganisms could provide new sources of these enzymes, with different specificities and better activities. In this context, we have improved lipase production by Luteibacter sp. by using basal medium supplemented with 2 % olive oil, a pH of 6 and a growth temperature of 37 °C. The enzyme extraction process with the addition of 0.25 % Tween 80 increased lipase activity. Implementation of these modifications increased lipase activity by approximately 430 %. The lipase activities produced in the culture supernatant (LCS) and extracted with Tween 80 (LCST80) were characterized. Both extracts hydrolyzed ρ-nitrophenyl (ρNP) esters with different acyl chain lengths, with a preference for short acyl lengths, and had optimum activity at 45 °C. The LCS was stable at acidic and alkaline pH, but LCST80 was only stable at alkaline pH. Methanol, SDS, Triton X-100, EDTA, and EGTA did not affect lipase activity, while divalent cations (Ca2+, Zn2+, Mg2+) - with the exception of Co2+— increased lipase activity. Both extracts showed transesterification activity on ρNP ester substrates, and both were able to hydrolyze different natural lipids. The characterization of lipase produced by Luteibacter sp. introduces this recently described genus as a new source of lipases with great biotechnological potential.  相似文献   

13.
Pseudomonas sp. lipase was immobilized onto glutaraldehyde-activated Florisil® support via Schiff base formation and stabilized by reducing Schiff base with sodium cyanoborohydride. The immobilization performance was evaluated in terms of bound protein per gram of support (%) and recovered activity (%). A 4-factor and 3-level Box–Behnken design was applied for the acylation of (±)-2-(propylamino)-1-phenylethanol, a model substrate, with vinyl acetate and the asymmetric acylations of other (±)-2-amino-1-phenylethanols with different alkyl substituents onto nitrogen atom such as (±)-2-(methylamino)-1-phenylethanol, (±)-2-(ethylamino)-1-phenylethanol, (±)-2-(butylamino)-1-phenylethanol and (±)-2-(hexylamino)-1-phenylethanol were performed under the optimized conditions. The optimal conditions were bulk water content of 1.8%, reaction temperature of 51.5 °C, initial molar ratio of vinyl acetate to amino alcohol of 1.92, and immobilized lipase loading of 47 mg mL?1. (R)-enantiomers of tested amino alcohols were preferentially acylated and the reaction purely took place on the hydroxyl group of 2-amino-1-phenylethanols. The increase of alkyl chain length substituted onto nitrogen atom caused an increase in the acylation yield and ee values of (S)-enantiomers. Enantiomeric ratio values were >200 for all the reactions. Our results demonstrate that the immobilized lipase is a promising biocatalyst for the preparation of (S)-2-amino-1-phenylethanols and their corresponding (R)-esters via O-selective acylation of (±)-2-amino-1-phenylethanols with vinyl acetate.  相似文献   

14.
Two diastereoisomers, 5R,6R-5-hydroxy-6(9α)-oxido-11α,15S-dihydroxyprost-13-enoic acid (7) and 5S,6S-5-hydroxy-6(9α)-oxido-11α,15S-dihydroxyprost-13-enoic acid (10) were synthesized for evaluation as possible biosynthetic intermediates in the enzymatic transformation of PGH2 or PGG2 into PGI2. The synthetic sequence entails the stereospecific reduction of the 9-keto function in PGE2 methyl ester after protecting the C-11 and C-15 hydroxyls as tbutyldimethylsilyl ethers. The resulting PGF derivative was epoxidized exclusively at the C-5 (6) double bond to yield a mixture of epoxides, which underwent facile rearrangement with SiO2 to yield the 5S,6S and 5R,6R-5-hydroxy-6(9α)-oxido cyclic ethers. It was found that dog aortic microsomes were unable to transform radioactive 9β-5S,6S[3H] or 9β-5R,6R[3H]-5-hydroxy-6(9α)-oxido cyclic ethers into PGI2. Also, when either diastereoisomer was included in the incubation mixture, neither isomer diluted the conversion of [1-14C]arachidonic acid into [1-14C]PGI2.  相似文献   

15.
Abstract: [(2S,2′R,3′R)-2-(2′,3′-[3H]Dicarboxycyclopropyl)glycine ([3H]DCG IV) binding was characterized in vitro in rat brain cortex homogenates and rat brain sections. In cortex homogenates, the binding was saturable and the saturation isotherm indicated the presence of a single binding site with a KD value of 180 ± 33 nM and a Bmax of 780 ± 70 fmol/mg of protein. The nonspecific binding, measured using 100 µM LY354740, was <30%. NMDA, AMPA, kainate, l (?)-threo-3-hydroxyaspartic acid, and (S)-3,5-dihydroxyphenylglycine were all inactive in [3H]DCG IV binding up to 1 mM. However, several compounds inhibited [3H]DCG IV binding in a concentration-dependent manner with the following rank order of potency: LY341495 = LY354740 > DCG IV = (2S,1′S,2′S)-2-(2-carboxycyclopropyl)glycine > (1S,3R)-1-aminocyclopentane-1,3-dicarboxylic acid > (2S,1′S,2′S)-2-methyl-2-(2-carboxycyclopropyl)glycine > l -glutamate = ibotenate > quisqualate > (RS)-α-methyl-4-phosphonophenylglycine = l (+)-2-amino-3-phosphonopropionic acid > (S)-α-methyl-4-carboxyphenylglycine > (2S)-α-ethylglutamic acid > l (+)-2-amino-4-phosphonobutyric acid. N-Acetyl-l -aspartyl-l -glutamic acid inhibited the binding in a biphasic manner with an IC50 of 0.2 µM for the high-affinity component. The binding was also affected by GTPγS, reducing agents, and CdCl2. In parasagittal sections of rat brain, a high density of specific binding was observed in the accessory olfactory bulb, cortical regions (layers 1, 3, and 4 > 2, 5, and 6), caudate putamen, molecular layers of the hippocampus and dentate gyrus, subiculum, presubiculum, retrosplenial cortex, anteroventral thalamic nuclei, and cerebellar granular layer, reflecting its preferential (perhaps not exclusive) affinity for pre- and postsynaptic metabotropic glutamate mGlu2 receptors. Thus, the pharmacology, tissue distribution, and sensitivity to GTPγS show that [3H]DCG IV binding is probably to group II metabotropic glutamate receptors in rat brain.  相似文献   

16.
Bestatin, [(2S,3R)-3-amino-2-hydroxy-4-phenylbutanoyl]-(S)-leucine, inhibited aminopeptidase B and leucine aminopeptidase in a competitive manner and their Ki values were calculated to be 6 × 10?8 and 2 × 10?8M, respectively. Among all stereoisomers of bestatin synthesized, those which have a (2S)-configuration in the 3-amino-2-hydroxy-4-phenylbutanoyl moiety showed marked inhibition against aminopeptidase B and leucine aminopeptidase compared with the other isomers which have (2R)-configuration. One of the isomers, [(2S,3S)-3-amino-2-hydroxy-4-phenylbutanoyl]-(R)-leucine, showed somewhat stronger activity against aminopeptidase B than bestatin. Aminopeptidase B appears to be a metallo-exopeptidase. It is proposed that bestatin and its active isomers are effective due to a mechanism other than a chelating action at the active center.  相似文献   

17.
This study provides the first measurements of the standard respiration rate (RS) and growth dynamics of European sardine Sardina pilchardus larvae reared in the laboratory. At 15° C, the relationship between RS (µl O2 individual?1 h?1) and larval dry mass (MD, µg) was equal to: RS = 0·0057(±0·0007, ± s.e.)·MD0·8835(±0·0268), (8–11% MD day?1). Interindividual differences in RS were not related to interindividual differences in growth rate or somatic (Fulton's condition factor) or biochemical‐based condition (RNA:DNA).  相似文献   

18.
Leaf respiration (R L) of evergreen species co-occurring in the Mediterranean maquis developing along the Latium coast was analyzed. The results on the whole showed that the considered evergreen species had the same R L trend during the year, with the lowest rates [0.83 ± 0.43 μmol(CO2) m?2 s?1, mean value of the considered species] in winter, in response to low air temperatures. Higher R L were reached in spring [2.44 ± 1.00 μmol(CO2) m?2 s?1, mean value] during the favorable period, and in summer [3.17 ± 0.89 μmol(CO2) m?2 s?1] during drought. The results of the regression analysis showed that 42% of R L variations depended on mean air temperature and 13% on total monthly rainfall. Among the considered species, C. incanus, was characterized by the highest R L in drought [4.93 ± 0.27 μmol(CO2) m?2 s?1], low leaf water potential at predawn (Ψpd= ?1.08 ± 0.18 MPa) and midday (Ψmd = ?2.75 ± 0.11 MPa) and low relative water content at predawn (RWCpd = 80.5 ± 3.4%) and midday (RWCmd = 67.1 ± 4.6%). Compared to C. incanus, the sclerophyllous species (Q. ilex, P. latifolia, P. lentiscus, A. unedo) and the liana (S. aspera), had lower R L [2.72 ± 0.66 μmol(CO2) m?2 s?1, mean value of the considered species], higher RWCpd (91.8 ± 1.8%), RWCmd (82.4 ± 3.2%), Ψpd (?0.65 ± 0.28 MPa) and Ψmd (?2.85 ± 1.20 MPa) in drought. The narrow-leaved species (E. multiflora, R. officinalis, and E. arborea) were in the middle. The coefficients, proportional to the respiration increase for each 10°C rise (Q10), ranging from 1.49 (E. arborea) to 1.98 (A. unedo) were indicative of the different sensitivities of the considered species to air temperature variation.  相似文献   

19.
The lipase secreted by Burkholderia cepacia ATCC 25416 was particularly attractive in detergent and leather industry due to its specific characteristics of high alkaline and thermal stability. The lipase gene (lipA), lipase chaperone gene (lipB), and native promoter upstream of lipA were cloned. The lipA was composed of 1095 bp, corresponding to 364 amino acid residues. The lipB located immediately downstream of lipA was composed of 1035 bp, corresponding to 344 amino acid residues. The lipase operon was inserted into broad host vector pBBRMCS1 and electroporated into original strain. The homologous expression of recombinant strain showed a significant increase in the lipase activity. LipA was purified by three-step procedure of ammonium sulfate precipitation, phenyl-sepharose FF and DEAE-sepharose FF. SDS-PAGE showed the molecular mass of the lipase was 33 kDa. The enzyme optimal temperature and pH were 60 °C and 11.0, respectively. The enzyme was stable at 30–70 °C. After incubated in 70 °C for 1 h, enzyme remained 72% of its maximal activity. The enzyme exhibited a good stability at pH 9.0–11.5. The lipase preferentially hydrolyzed medium-chain fatty acid esters. The enzyme was strongly activated by Mg2+, Ca2+, Cu2+, Zn2+, Co2+, and apparently inhibited by PMSF, EDTA and also DTT with SDS. The enzyme was compatible with various ionic and non-ionic surfactants as well as oxidant H2O2. The enzyme had good stability in the low- and non-polar solvents.  相似文献   

20.
《Process Biochemistry》2007,42(9):1312-1318
A combination of the enzymatic resolution and chemical racemization for the heterogeneous sequential kinetic resolution (SKR) was employed to resolve (R,S)-2-octanol under microwave irradiation. Mesoporous molecular sieves SBA-15, alumina and strong basic styrene anion exchange resin were screened and selected as the optimum supports to immobilize the lipase from Pseudomonas sp. (PSL), oxidant-Chromium trioxide (CrO3) and reductant-Sodium borohydride (NaBH4), respectively. The immobilized catalysts exhibited good reusability: it remained 90%, 72% and 80% of their initial activities after five reuses for the immobilized lipase, the immobilized oxidant and the immobilized reductant, respectively. Further, the E values of the immobilized PSL was increased from 23 under conventional heating to 40 under microwave irradiation in resolution of (R,S)-2-octanol. The immobilized catalysts were then used in SKR of (R,S)-2-octanol under microwave irradiation after optimizing the reaction media. Under the optimum conditions, (R)-2-octanol acetate was obtained at 99% enantiomeric excess with 84% yield in 2 h.  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