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 15 毫秒
1.
Partially purified Trichoderma reesei RUT-C30 acetyl esterase preparation was found to catalyze acyl transfer reactions in organic solvents, mixtures of organic solvents with water and even in water. Using different acyl donors, the best results for acetyl transfer in water were obtained using vinyl acetate. As acetyl acceptors, a variety of hydroxyl bearing compounds in aqueous solutions were used. Degree of conversion and the number of newly formed acetates varied according to the acceptor used. Conversions over 50% were observed for the majority of several common monosaccharides, their methyl and deoxy derivatives and oligosaccharides. In several cases, the transesterification reaction exhibited strict regioselectivity, leading to only one acetyl derivative. Preparative potential of the transesterification in water was demonstrated by acetylation of methyl β- -glucopyranoside, 4-nitrophenyl β- -glucopyranoside and kojic acid, yielding 56.4% of methyl 3-O-acetyl β- -glucopyranoside, 70.2% of 4-nitrophenyl 3-O-acetyl β- -glucopyranoside and 30.9% of 7-O-acetyl-kojic acid as the only reaction products.

This enzymatically catalyzed transacetylation in water, which is applied to transformation of saccharides for the first time, opens a new area in chemoenzymatic synthesis. Its major advantages are simplicity, highly regioselective esterification of polar compounds, high yields, low enzyme consumption and elimination of the need to use toxic organic solvents.  相似文献   


2.
Zhuo K  Liu H  Zhang X  Liu Y  Wang J 《Carbohydrate research》2008,343(14):2428-2432
The (13)C NMR spectra of methyl beta-d-glucopyranoside, methyl beta-d-galactopyranoside, methyl beta-d-xylopyranoside, and methyl beta-l-arabinopyranoside were recorded in CaCl(2)/KCl+D(2)O mixtures and in D(2)O. The chemical shifts of C-1, C-3, and C-5 in the methyl beta-d-glucopyranoside and methyl beta-d-galactopyranoside decrease rapidly as molalities of CaCl(2)/KCl increase, while those of C-1, C-2, and C-3 in the methyl beta-d-xylopyranoside and methyl beta-l-arabinopyranoside decrease rapidly as molalities of CaCl(2)/KCl increase. Cations (Ca(2+)/K(+)) can weakly complex with O in OMe of the pyranosides studied. Results are discussed in terms of the stereochemistry of the pyranoside molecules and the structural properties of the ions.  相似文献   

3.
The objective of this work is to assess the structure and activity of Candida rugosa lipase (CRL) pretreated with seventeen ionic liquids (ILs), five organic solvents and super-critical carbon dioxide (SC-CO2). The results revealed that anion selection of ILs showed generally much greater effects on CRL esterification activity than cation choice, and CRL pretreated by ILs with strong water miscible properties showed very low esterification activity. The highest CRL activity treated with ILs [Hmim][PF6] was obtained with the value of 45078.0 U/g-protein. Furthermore, the CRL activities pretreated with five conventional organic solvents were also examined and the values increased with the log P decrease of organic solvents when log P was lower than 2.0. Finally, the CRL activities were respectively 1.2- and 1.3-fold higher over the untreated ones after pretreatment with sub- and super-critical CO2 under the pressures of 6 MPa and 15 MPa at 40 °C for 20 min. Further analyses via FT-IR demonstrated that the high activity of CRL pretreated with ILs, organic solvents and SC-CO2 was probably caused by the changes of CRL secondary structure. In conclusion, the results in this work will be helpful for us to choose the suitable reaction medium in CRL biocatalysis and biotransformation reactions.  相似文献   

4.
An interesting observation was found during our continued studies on the hydrolysis of ibuprofen esters by Candida rugosa lipase (CRL). An important role is played by pH in the stereospecific hydrolysis of these esters. The flap region of CRL plays a significant role in the access of the substrate to the active site of the enzyme. At pH 5.6, 48% of the methyl ester and 5% of the butyl ester of ibuprofen were hydrolysed in 5.5 h, whereas at pH 7.2, 9% of methyl ester and 45% of the butyl ester of ibuprofen was hydrolysed in a identical reaction time using CRL. This lead us to assume that CRL prefers the methyl ester of ibuprofen as a substrate at an acidic pH and the butyl ester of ibuprofen at a neutral pH. Therefore, in order to understand the role of pH in the substrate selection by CRL for the esters of ibuprofen we used the crystallographic coordinates of the open form of the CRL (1CRL) for molecular dynamics (MD) simulations under acidic and neutral conditions for 2 ns using GROMACS. The final structures obtained after simulation in acidic and neutral conditions were compared with the energy-minimized structure, and the root-mean-square deviations (r.m.s.ds) were calculated. The r.m.s.d. of the CRL flap at neutral pH was found to be greater than that of the CRL flap at acidic pH. The extent to which the flap opens at neutral pH allowed the bulkier substrate, the butyl ester of ibuprofen, to diffuse into the active site and provides the best enzyme-substrate fit for this specific substrate. At acidic pH there is a decreased opening of the flap thereby accommodating a more compact substrate, namely the methyl ester of ibuprofen. Thus, simulation experiments using MD provide reasonable insight for the pH-dependent substrate selectivity of CRL in aqueous environments.  相似文献   

5.
Summary Regioselective ethanolysis of peracylated methyl , -D-glucopyranoside and methyl -D-mannopyranoside in anhydrous organic solvent (n-hexane/EtOH = 99/1) could afford 6-OH derivatives exclusively by Candida rugosa lipase (CRL). No 4 6 acyl migration was observed in such an anhydrous solvent system. Substrates with propanoyl groups were more reactive than with acetyl groups on CRL-catalyzed reactions.  相似文献   

6.
Crich D  Yao Q  Bowers AA 《Carbohydrate research》2006,341(10):1748-1752
The N-bromosuccinimide mediated fragmentation of methyl 4,6-O-benzylidene-beta-D-galactopyranoside results in the formation of methyl 4-O-benzoyl-6-bromo-6-deoxy-beta-D-galactopyranoside and methyl 4-O-benzoyl-3-bromo-3-deoxy-beta-D-gulopyranoside, as opposed to the methyl 6-O-benzoyl-3-bromo-3-deoxy-beta-D-gulopyranoside originally reported. The kinetic methyl 4-O-benzoyl-6-bromo-6-deoxy-beta-D-galactopyranoside rearranges to the thermodynamic methyl 4-O-benzoyl-3-bromo-3-deoxy-beta-D-gulopyranoside under the reaction conditions, likely via a 3,6-anhydro galactopyranoside. The NBS-mediated cleavage of 4,6-O-benzylidene acetals in the galactopyranoside series is therefore shown to conform to the regiochemistry observed in the corresponding gluco- and mannopyranoside series with preferential cleavage of the C6-O6 bond by an ionic mechanism.  相似文献   

7.
Homogalacturonan-derived partly methylated and/or acetylated oligogalacturonates were recovered after enzymatic hydrolysis (endo-polygalacturonase+pectin methyl esterase+side-chain degrading enzymes) of sugar beet pectin followed by anion-exchange and size exclusion chromatography. Around 90% of the GalA and 75% of the acetyl groups present in the initial sugar beet pectin were recovered as homogalacturonan-derived oligogalacturonates, the remaining GalA and acetyl belonging to rhamnogalacturonic regions. Around 50% of the acetyl groups present in sugar beet homogalacturonans were recovered as partly methylated and/or acetylated oligogalacturonates of degree of polymerisation 5 whose structures were determined by electrospray ionization ion trap mass spectrometry (ESI-IT-MSn). 2-O-acetyl- and 3-O-acetyl-GalA were detected in roughly similar amounts but 2,3-di-O-acetylation was absent. Methyl-esterified GalA residues occurred mainly upstream 2-O-acetyl GalA. Oligogalacturonates containing GalA residues that are at once methyl- and acetyl-esterified were recovered in very limited amounts. A tentative mapping of the distribution of acetyl and methyl esters within sugar beet homogalacturonans is proposed. Unsubstituted GalA residues are likely to be present in limited amounts (approximately 10% of total GalA residues), due to the fact that methyl and acetyl groups are assumed to be most often not carried by the same residues.  相似文献   

8.
Regioselective synthesis of L-phenylalanyl ester of D-glucose with unprotected L-phenylalanine and D-glucose was carried out in organic solvents using lipases from Rhizomucor miehei (RML) and porcine pancreas (PPL). The reaction was investigated in terms of free unprotected L-phenylalanine, D-glucose, RML and PPL concentrations, buffer salts (pH and concentration), enzyme reusability and incubation period. Under the experimental conditions employed, both the enzymes exhibited good esterification potentialities, with RML exhibiting better conversions (maximum yield, 98%) than PPL (maximum yield, 75.6%). Reactions in the presence of buffer salts gave about 10% higher yields than those in their absence. Two-dimensional heteronuclear single quantum coherence transfer (HSQCT) NMR spectral analysis confirmed the formation of five diastereomeric esters: three different L-phenylalanyl-D-glucose monoesters (6-O: 24.1%, 3-O: 23.3% and 2-O: 19.2%) and two different diesters (2,6-di-O: 16.6% and 3,6-di-O: 16.8%) in an esterification yield of 92.4%.  相似文献   

9.
The rates of hydrolysis of acetyl phosphate in the presence of 0.1 M NaOH and of ATP in the presence of either 1 M HCl or 1 M NaOH were measured at different temperatures and in the presence of different concentrations of the organic solvents dimethyl sulfoxide or ethylene glycol. Under all conditions tested, there was a progressive increase in the rate constant of hydrolysis of both phosphate compounds as the water activity of the medium was decreased by the addition of organic solvents. At 25 degrees C, substitution of 70% of the water of the medium by dimethyl sulfoxide promoted an increase of two orders of magnitude in the rate constant of acetyl phosphate hydrolysis. In the presence of 80% and 90% dimethyl sulfoxide the rate of acetyl phosphate hydrolysis increased by more than two orders of magnitude and was so fast that it could not be measured with the method used. The effect of organic solvents on the rate of ATP hydrolysis was less pronounced than that observed for acetyl phosphate hydrolysis. At 30 degrees C, substitution of 90% of water by an organic solvent promoted a 4-6-fold increase of the rate of ATP hydrolysis. Acceleration of either acetyl phosphate or ATP hydrolysis rates was promoted by a decrease in both activation energies (Ea) and in entropies of activation delta S. The data obtained are discussed with reference to the mechanism of catalysis of enzymes involved in energy transduction such as the Ca2+-ATPase of sarcoplasmic reticulum and the F1-ATPase of mitochondria.  相似文献   

10.
D Yang  V Kumar 《Carbohydrate polymers》2012,90(4):1486-1493
In this paper, we report the preparation of oxidized cellulose acetate methyl esters (OCAM) from OCA (OC14A: carboxylic acid content 10.6% (w/w), degree of acetyl group substitution: 1.89; OC21A: carboxylic acid content 15.7% (w/w), degree of acetyl group substitution: 1.70) by treatment with methanol at room temperature using 4-dimethylaminopyridine (DMAP) as a catalyst and dicyclohexylcarbodiimide (DCC) as a coupling agent. The new polymers were characterized by Fourier-transform infrared (FT-IR) and (1)H and (13)C nuclear magnetic resonance spectroscopies, carboxylic acid content determination, moisture sorption isotherms, intrinsic viscosity, and powder X-ray diffractometry. The new polymers are amorphous powders. It is practically insoluble in water but show solubility in a range of organic solvents.  相似文献   

11.
Within a research project aimed at probing the substrate specificity and the enantioselectivity of Candida rugosa lipase (CRL), computer modeling studies of the interactions between CRL and methyl (+/-)-2-(3-benzoylphenyl)propionate (Ketoprofen methyl ester) have been carried out in order to identify which amino acids are essential to the enzyme/substrate interaction. Different binding models of the substrate enantiomers to the active site of CRL were investigated by applying a computational protocol based on molecular docking, conformational analysis, and energy minimization procedures. The structural models of the computer generated complexes between CRL and the substrates enabled us to propose that Phe344 and Phe345, in addition to the residues constituting the catalytic triad and the oxyanion hole, are the amino acids mainly involved in the enzyme-ligand interactions. To test the importance of these residues for the enzymatic activity, site-directed mutagenesis of the selected amino acids has been performed, and the mutated enzymes have been evaluated for their conversion and selectivity capabilities toward different substrates. The experimental results obtained in these biotransformation reactions indicate that Phe344 and especially Phe345 influence CRL activity, supporting the findings of our theoretical simulations.  相似文献   

12.
Chemical modification of lysine residues in Candida rugosa lipase (CRL) was carried out using five different functional ionic liquids, and about 15.4–25.0 % of the primary amino groups of lysine were modified. Enzymatic properties of the native and modified CRLs were investigated in olive oil hydrolysis reaction. Improved thermal stability, catalytic activity in organic solvents, and adaptability to temperature and pH changes were achieved compared with the native enzyme. CRL modified by [choline][H2PO4] showed the best results, bearing a maximum improvement of 16.7 % in terms of relative activity, 5.2-fold increase in thermostability (after incubation at 45 °C for 5 h), and 2.3-fold increase in activity in strong polar organic solvent (80 % dimethyl sulfoxide) compared with the native enzyme. The results of ultraviolet, circular dichroism and fluorescence spectroscopy suggested that the change of the secondary and tertiary structures of CRL caused by the chemical modification resulted in the enhancement of enzymatic performance. The modification of CRL with functional ionic liquids was proved to be a novel and efficient method for improving the enzymatic properties of CRL.  相似文献   

13.
S L Ginell  S Kuzmich  R A Jones  H M Berman 《Biochemistry》1990,29(46):10461-10465
The crystal and molecular structure of the first DNA duplex containing the carcinogenic lesion O6MeG has been determined to a resolution of 1.9 A and refined to an R factor of 19%. (d[CGC-(O6Me)GCG])2 crystallizes in the left-handed Z DNA form and has crystal parameters and conformational features similar to those of the parent sequence [d(CG)3]2. The methyl groups on O6 of G4 and G10 have C5-C6-O6-O6Me torsion angles of 73 degrees and 56 degrees, respectively, and protrude onto the major groove surface. The base-pairing conformation for the methylated G.C base pairs is of the Watson-Crick type as opposed to a wobble-type conformation that had been proposed in a B DNA fragment. As in other Z DNA structures, a spine of hydration is seen in the minor groove.  相似文献   

14.
Previously, it was demonstrated that pancreatic cholesterol esterase is selectively inhibited by 6-chloro-2-pyrones with cyclic aliphatic substituents in the 3-position. Inhibition is reversible and is competitive with substrate. Pancreatic cholesterol esterase is a potential target for treatment of hypercholesterolemia. In the present study, yeast cholesterol esterase from Candida cylindracea (also called C. rugosa CRL3) was compared to porcine pancreatic cholesterol esterase for inhibition by a series of 3-alkyl- or 5-alkyl-6-chloro-2-pyrones. In addition, CRL3 was compared with the related yeast lipase CRL1. Inhibition of CRL3 by substituted 6-chloro-2-pyrones was competitive with binding of the substrate p-nitrophenyl butyrate. Inhibition constants ranged from 0.2 microM to >90 microM. Small changes in the alkyl group had profound effects on binding. The pattern of inhibition of CRL3 is quite distinct from that observed with porcine cholesterol esterase. Molecular modeling studies suggest that the orientation of binding of these inhibitors at the active site of CRL3 can vary but that the pyrone ring consistently occupies a position close to the active site serine. CRL1 is highly homologous to CRL3. Nevertheless, patterns of inhibition of CRL1 by substituted 6-chloro-2-pyrones differ markedly from patterns observed with CRL3. The substituted 6-chloro-2-pyrones are slowly hydrolyzed in the presence of CRL1 and are pseudosubstrates of CRL3, but are simple reversible inhibitors of pancreatic cholesterol esterase  相似文献   

15.
Peracetylated samples of cyclomaltoheptaose (β-cyclodextrin, β-CD) and its methylated derivative were studied by 13C NMR. The acetyl carbonyl carbon signal in peracetylated β-CD was resolved into a triplet, and the three peaks were assigned by long-range C---H COSY and INAPT techniques. The individual carbonyl peak was found to be indicative of the location of the acetyl group on the 2, 3, and 6 position in the glucose residues. An acetylated derivative of a partly methylated β-CD was also subjected to 13C NMR analysis to determine the distribution of acetyl and, subsequently, methyl groups on the glucose residues.  相似文献   

16.
We used substrate mapping to develop a rule that predicts which enantiomer of chiral carboxylic acid esters reacts faster in hydrolyses catalyzed by lipase from Candida rugosa (CRL, triacylglycerol hydrolase, E. C. 3.1.1.3). This rule, based on the size of the substituents at the stereocenter, is not reliable for crude CRL. It predicts the favoured enantiomer for only 23 out of 34 examples, 68% reliability. However, this rule is completely reliable for purified CRL; it predicts the favoured enantiomer for all 16 examples correctly. The examples include arylpropanoicacids, aryloxypropanoic acids, α-halophenylacetic acids, mandelic acid and O-methylmandelic acid. Further, purified CRL did not catalyse the hydrolysis of N-CBZ-phenylalanine methyl ester and N-CBZ-norleucine methyl ester. These two substrates were exceptions to the rule with crude CRL as the catalyst. Besides eliminating several exceptions, purification also raised the enantioselectivity of CRL toward carboxylic acid esters. To provide a structural basis for this proposed rule we examined the x-ray crystal structure of CRL containing transition state analogs of ester hydrolysis. We suggest that the large substituent of chiral carboxylic acids binds in a tunnel that normally binds the alkyl chain of a fatty acid. The phenyl rings of Phe 345 and Phe 415 lie close to the stereocenter, thereby fixing the orientation of the medium substituent. The three-dimensional orientation of these proposed binding sites is consistent with the rule derived from substrate mapping.  相似文献   

17.
Lipases from two different sources Candida rugosa (CRL) and Burkholderia cepacia (BCL) were formulated as enzyme precipitated and rinsed with organic solvents, organic solvent rinsed enzyme preparation, cross-linked enzyme aggregates (CLEAs) and protein coated micro-crystals (PCMCs). These various enzyme formulates were evaluated for the kinetic resolution of (+/-)-1-phenylethanol in ionic liquid [Bmim][PF(6)] by transesterification with vinyl acetate. Of all the enzyme forms evaluated EPRP and PCMC in the case of CRL showed the best results with 26 % (E value=153) and 53% (E value=79) conversion, respectively, at 35 degrees C in 24h. Carrying out this conversion with PCMC at lower temperature of 25 degrees C further improved the E value to 453 (with 44% conversion in 12h). For BCL the acetone-rinsed enzyme preparation (AREP), CLEA and PCMC performed equally well with % conversion of 50 and 99 ee(p) (%) (E value >1000) in just 2h, whereas, the free lipase gave only 8% conversion.  相似文献   

18.
The positions of esterification of the 4 to 5 acetyl residues in the acetylated methylmannose-containing polysaccharide from Streptomyces griseus have been established by the methyl replacement technique, wherein ester substituents are specifically replaced with methyl ether substituents. The newly incorporated methyl groups were distinguished from 3-O-methyl groups by the use of polysaccharide containing radioactively labeled endogenous methyl groups. The positions of methyl group localization were established by a proton magnetic resonance study of the intact methyl-replaced polysaccharide combined with an analysis of the constituent monosaccharides by gas-liquid chromatography-electron impact mass spectrometry of their alditol acetate derivatives. These studies demonstrate that the acetyl groups are located at position 6 of approximately half of the 10 contiguous alpha(1 leads to 4)-linked 3-O-methyl-D-mannose residues. Purification of the polysaccharide was accomplished by an added step involving affinity chromatography on a column containing immobilized palmitoyl residues. The affinity of the polysaccharide for this long chain lipid suggests that its plays a role similar to the methylmannose-containing polysaccharide of Mycobacterium smegmatis in its regulation of the bacterium's fatty acid synthetase.  相似文献   

19.
Thirteen glucose analogues bearing electrophilic groups were synthesized (five of them for the first time) and screened as inhibitors of the glucose transporter (EIIGlc) of the Escherichia coli phosphoenolpyruvate-sugar phosphotransferase system (PTS). 2',3'-Epoxypropyl beta-d-glucopyranoside (3a) is an inhibitor and also a pseudosubstrate. Five analogues are inhibitors of nonvectorial Glc phosphorylation by EIIGlc but not pseudosubstrates. They are selective for EIIGlc as demonstrated by comparison with EIIMan, another Glc-specific but structurally different transporter. 3a is the only analogue that inhibits EIIGlc by binding to the high-affinity cytoplasmic binding site and also strongly inhibits sugar uptake mediated by this transporter. The most potent inhibitor in vitro, methyl 6,7-anhydro-d,l-glycero-alpha-d-gluco-heptopyranoside (1d), preferentially interacts with the low-affinity cytoplasmic site but only weakly inhibits Glc uptake. Binding and/or phosphorylation from the cytoplasmic side of EIIGlc is more permissive than sugar binding and/or translocation of substrates via the periplasmic site. EIIGlc is rapidly inactivated by the 6-O-bromoacetyl esters of methyl alpha-d-glucopyranoside (1a) and methyl alpha-d-mannopyranoside (1c), methyl 6-deoxy-6-isothiocyanato-alpha-d-glucopyranoside (1e), beta-d-glucopyranosyl isothiocyanate (3c) and beta-d-glucopyranosyl phenyl isothiocyanate (3d). Phosphorylation of EIIGlc protects, indicating that inactivation occurs by alkylation of Cys421. Glc does not protect, but sensitizes EIIGlc for inactivation by 1e and 3d, which is interpreted as the effect of glucose-induced conformational changes in the dimeric transporter. Glc also sensitizes EIIGlc for inactivation by 1a and 1c of uptake by starved cells. This indicates that Cys421 which is located on the cytoplasmic domain of EIIGlc becomes transiently accessible to substrate analogues on the periplasmic side of the transporter.  相似文献   

20.
《Process Biochemistry》2007,42(6):1021-1027
Candida rugosa lipase (CRL) was immobilized on Amberlite XAD 7 and the advantage of immobilization under the best reaction conditions in achieving high activity and enantioselectivity was shown for the hydrolysis of racemic Naproxen methyl ester. The performance of CRL was found to be better when the enzyme was immobilized at the temperature and pH values where higher conversion and enantioselectivity were obtained. The effects of immobilized lipase load, temperature, pH and substrate concentration on the conversion and enantioselectivity toward S-Naproxen production in aqueous phase/isooctane biphasic batch system were also evaluated. The increase in immobilized lipase load in 320–800 U/mL range increased the conversion of the substrate and enantioselectivity for S-Naproxen. The kinetic resolution of racemic Naproxen methyl ester conducted at the temperatures of 40, 45 and 50 °C and at the pH values of 4, 6, 7.5 and 9 resulted in the highest conversion and enantioselectivity at 45 °C and pH 6. Higher concentration of racemic Naproxen methyl ester than 10 mg/mL decreased both the conversion and enantioselectivity. CRL, which was immobilized at the temperature and pH values where the enzyme was more enantioselective, was successfully used in three successive batch runs each of 180 h. The highest enantiomeric ratio achieved in the S-Naproxen production was 174.2 with the conversion of 49%.  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