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 11 毫秒
1.
Candida antarctica lipase B catalyzed the stereoselective ammoniolysis of N‐alkyl aziridine‐2‐carboxylates in tBuOH saturated with ammonia and yielded the (2S)‐aziridine‐2‐carboxamide and unreacted (2R)‐aziridine‐2‐carboxylate. Varying the N‐1 substituent on the aziridine ring changed the rate and stereoselectivity of the reaction. Substrates with a benzyl substituent or a (1′R)‐1‐phenylethyl substituent reacted approximately ten times faster than substrates with a (1′S)‐1‐phenylethyl substituent. Substrates with a benzyl substituent showed little stereoselectivity (E=5–7) while substrates with either a (1′R)‐ or (1′S)‐1‐phenylethyl substituent showed high stereoselectivity (D>50). Molecular modeling by using the current paradigm for enantioselectivity—binding of the slow enantiomer by an exchange‐of‐substituents orientation—could not account for the experimental results. However, modeling an umbrella‐like‐inversion orientation for the slow enantiomer could account for the experimental results. Steric hindrance between the methyl in the (1′S)‐1‐phenylethyl substituent and Thr138 and Ile189 in the acyl‐binding site likely accounts for the slow reaction. Enantioselectivity likely stems from an unfavorable interaction of the methine hydrogen with Thr40 for the slow enantiomer and from subtle differences in the orientations of the other three substituents. This success in rationalizing the enantioselectivity supports the notion that an umbrella‐like‐inversion orientation can contribute to enantioselectivity in lipases.  相似文献   

2.
Both enantiomers of optically pure 4‐bromo‐3‐hydroxybutanoate, which is an important chiral building block in the syntheses of various biologically active compounds including statins, were synthesized from rac‐4‐bromomethyl‐β‐lactone through kinetic resolution. Candida antarctica lipase B (CAL‐B) enantioselectively catalyzes the ring opening of the β‐lactone with ethanol to yield ethyl (R)‐4‐bromo‐3‐hydroxybutanoate with high enantioselectivity (E>200). The unreacted (S)‐4‐bromomethyl‐β‐lactone was converted to ethyl (S)‐4‐bromo‐3‐hydroxybutanoate (>99% ee), which can be further transformed to ethyl (R)‐4‐cyano‐3‐hydroxybutanoate, through an acid‐catalyzed ring opening in ethanol. Molecular modeling revealed that the stereocenter of the fast‐reacting enantiomer, (R)‐bromomethyl‐β‐lactone, is ∼2 Å from the reacting carbonyl carbon. In addition, the slow‐reacting enantiomer, (S)‐4‐bromomethyl‐β‐lactone, encounters steric hindrance between the bromo substituent and the side chain of the Leu278 residue, while the fast‐reacting enantiomer does not have any steric clash.  相似文献   

3.
Most lipases resolve secondary alcohols in accordance with the “Kazlauskas rule” to give the R enantiomers. In a similar manner to other lipases, Candida rugosa lipase (CRL) exhibits R enantioselectivity towards heptan‐2‐ol, although the enantiomeric ratio (E) is low (E=1.6). However, unexpected enantioselectivity (i.e., S enantioselectivity, E=58) of CRL towards 4‐(tert‐butoxycarbonylamino)butan‐2‐ol, which has a similar chain length to heptan‐2‐ol, has been observed. To develop a deeper understanding of the molecular basis for this unusual enantioselectivity, we have conducted a series of molecular modeling and substrate engineering experiments. The results of these computational and experimental analyses indicated that a hydrogen bond between the Ser450 residue and the nitrogen atom of the carbamate group is critical to stabilize the transition state of the S enantiomer.  相似文献   

4.
BACKGROUND: The enantiomers of N‐hydroxymethyl vince lactam are important intermediates during the synthesis of chiral drugs. The preparation of its single enantiomer can be performed through enzymatic resolution. The aim of this work is to obtain (1S, 4R)‐N‐hydroxymethyl vince lactam with high enantiomeric purity via lipase‐catalyzed enantioselective transesterification in organic solvents. To achieve this, effects of various reaction conditions (including lipase sources, acyl donor, substrate molar ratio, organic solvent, temperature, and water activity) on the enzyme activity as well as enantioselectivity were investigated. RESULTS: The results of the study showed that the enantiopreference for all the selected enzymes was (4S, 1R)‐N‐hydroxymethyl vince lactam in enantioselective transesterification of racemic N‐hydroxymethyl vince lactam. Under the selected optimum conditions, the highest enantioselectivity (E = 33.8) was obtained with a higher enzyme activity (20.3 µmol g?1 min?1) for Mucor miehei lipase (MML) when vinyl valerate was used as the acyl donor. Besides, the remained (1S, 4R)‐N‐hydroxymethyl vince lactam with high enantiomeric purity (ee > 99%) was obtained when the conversion was about 60%. CONCLUSION: The results obtained clearly demonstrated potential for industrial application of lipase in resolution of N‐hydroxymethyl vince lactam through enantioselective transesterification. © 2012 Society of Chemical Industry  相似文献   

5.
Lipase from Candida rugosa was immobilized on poly(N‐vinyl‐2‐pyrrolidone‐co‐styrene) hydrogel (poly‐(VP‐co‐ST)) with ethylene dimethacrylate and α,α'‐azoisobutyronitrile, which act as crosslinker and initiator, respectively. Three different compositions of monomers were used, namely VP(%):ST(%), 10:90, 50:50, and 70:30 (wt(%)/wt(%)). The immobilized lipases were used in the enantioselective esterification of (R,S)‐2‐(4‐chlorophenoxy)‐propanoic acid with n‐tetradecanol. The optimum reaction condition of the enantioselective esterification for the native lipase and the poly(VP‐co‐ST) hydrogel immobilized lipases was determined with respect to temperature, solvents, and initial water activity (aw). The optimum temperature obtained was 40°C, with the poly(VP‐co‐ST) hydrogel immobilized lipase VP(%)/ST(%):10:90 showing the highest enantiomeric excess. In the solvent effect studies, the best solvents for high enantioselectivity were chloroform and carbon tetrachloride. In the aw studies, optimum αw for NL, VP(%):ST(%), 10:90, and 50:50 was 0.328, while for VP(%):ST(%), 70:30, it was 0.55. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 92: 3381–3386, 2004  相似文献   

6.
Lipase (EC 3.1.1.3) was immobilized on cellulose acetate–TiO2 gel fibre by the sol–gel method. The immobilized lipases were used for esterification of n‐butyric acid with n‐butyl alcohol and enantioselective acylation of (R, S)‐phenylethanol using vinyl acetate as an acyl donor. Compared with native lipase, the activity of the immobilized lipase was stable and relatively unaffected by the water content of the solvent and the substrate concentration. The data indicate that the lipases are immobilized on the fibre surface and that enzyme activity is influenced by bound water. However, the thermal reactivity and enantioselectivity of the immobilized lipase were less than those of native lipase. This may not reflect thermal inactivation of the enzyme but rather significant thermal contraction of the gel fibre by cellulose crystallization, resulting in liberation of bound water and a decrease in the amount of enzyme which is available for the reaction. Copyright © 2001 Society of Chemical Industry  相似文献   

7.
The hydrolytic stability of a series of oligoesters comprised of three and four different monomers was evaluated. The hydroxyl terminal oligoesters were prepared from adipic acid (AA) and isophthalic acid (IPA), with six different diols and one triol, which included: 1,4‐butanediol, 1,5‐pentanediol, 1,6‐hexanediol, neopentyl glycol, 2‐methyl‐1,3‐propanediol, trimethylolpropane, and 2‐butene‐1,4‐diol. The hydroxyl terminated oligoesters were reacted with phenyl isocyanate to form telechelic urethane groups. Hydrolysis rate constants were obtained from plots of acid number vs. time. It was observed that ternary oligoester systems had lower hydrolysis rates than quaternary systems. In addition to investigating the hydrolytic stability of the synthesized oligoesters, polyurethane coatings were produced by reacting the hydroxyl‐terminated oligoesters with an aliphatic polyisocyanate (1,6‐hexanediisocyanate trimer). Model oliogester hydrolysis was then correlated to the weatherability of a crosslinked polyurethane film. © 2013 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2014 , 131, 40198.  相似文献   

8.
A lipase‐catalyzed enantioselective esterification process in cyclohexane was developed for the synthesis of S(+)‐naproxen ester prodrugs containing the moiety of N,N‐dialkylamino, ethylene glycol or alkyl ether of ethylene glycol. A high enantiomeric ratio of 44 was obtained when di(ethylene glycol) was selected as the best acyl acceptor. A reversible ping‐pong Bi Bi mechanism has been employed to elucidate the enzymatic behavior of the initial conversion rate for S(+)‐naproxen and the time‐course conversions for both enantiomers. Improvement of the enzyme activity was demonstrated when alcohol in excess of its cyclohexane solubility limit was used. The application of excess racemic naproxen in the presence of solid substrate suspensions showed enhanced productivity and enantioselectivity for the desired S(+)‐ester. Studies of the recovery and racemization of the remaining R(−)‐naproxen are also reported. © 1999 Society of Chemical Industry  相似文献   

9.
The hydrolytic dehalogenation of rac‐1,3‐dibromobutane catalyzed by the haloalkane dehalogenase LinB from Sphingobium japonicum UT26 proceeds in a sequential fashion: initial formation of intermediate haloalcohols followed by a second hydrolytic step to produce the final diol. Detailed investigation of the course of the reaction revealed favored nucleophilic displacement of the sec‐halogen in the first hydrolytic event with pronounced R enantioselectivity. The second hydrolysis step proceeded with a regioselectivity switch at the primary position, with preference for the S enantiomer. Because of complex competition between all eight possible reactions, intermediate haloalcohols formed with moderate to good ee ((S)‐4‐bromobutan‐2‐ol: up to 87 %). Similarly, (S)‐butane‐1,3‐diol was formed at a maximum ee of 35 % before full hydrolysis furnished the racemic diol product.  相似文献   

10.
The diastereomeric 1,4‐diphosphine ligands, (S,S,S,S)‐ 1a , (R,S,S,R)‐ 1b and (R,S,S,S)‐ 1c , with the imidazolidin‐2‐one backbone were synthesized, and utilized for an investigation of the effects of backbone chirality on the enantioselectivity in the Rh(I)‐catalyzed hydrogenation of various functionalized olefinic substrates. It was found that the catalytic efficiencies are largely dependent on the configurations of the α‐carbons to phosphine. Thus, the Rh complex of the pseudo‐C2‐symmetrical diphosphine, (R,S,S,S)‐ 1c , showed excellent enantioselectivities (93.0–98.6% ees) in the hydrogenations of a broad spectrum of substrates, and especially in the hydrogenations of methyl α‐(N‐acetyamino)‐β‐arylacrylates (95.3–97.0% ees). However, the enantioselectivities obtained with the C2‐symmetrical (R,S,S,R)‐ 1b were largely dependent on the substrate (19.8–97.3% ees). The Rh complex of ligand 1a having the (S,S,S,S)‐configuration showed the lowest catalytic efficiency for all of the substrates examined (0–84.8% ees).  相似文献   

11.
The Escherichia coli esterase YbfF displays high activity towards 1,2‐O‐isopropylideneglycerol (IPG) butyrate and IPG caprylate, and prefers the R‐enantiomer of these substrates, producing the S‐enantiomer of the IPG product in excess. To improve the potential of the enzyme for the kinetic resolution of racemic esters of IPG, an enhancement of the activity and enantioselectivity would be highly desirable. Molecular docking of the R‐enantiomer of both IPG esters into the active site of YbfF allowed the identification of proximal YbfF active site residues. Four residues (25, 124, 185 and 235) were selected as targets for mutagenesis, in order to enhance YbfF activity and enantioselectivity towards IPG esters. Random mutagenesis at positions 25, 124, 185 and 235 yielded several best YbfF variants with enhanced activity and enantioselectivity towards IPG esters. The best YbfF mutant, W235I, exhibited a 2‐fold higher enantioselectivity than wild‐type YbfF, with an E=38 for IPG butyrate and an E=77 for IPG caprylate. Molecular docking experiments further support the enhanced enantioselectivity shown experimentally and the structural effects of this amino acid substitution on the active site of YbfF are provided. The engineered W235I mutant is an attractive catalyst for practical applications in the kinetic resolution of IPG esters.  相似文献   

12.
The enantioselective esterification of (R,S)-ibuprofen has been performed with Novozym 435TM in solvent media under reduced pressure. The nature of the solvent affects the activity and enantioselectivity of the lipase. High esterification rates are obtained in solvents having high hydrophobicity (log P>4). On the other hand, the best enantioselectivity is obtained with solvents having low hydrophobicity (log P<3). This particular enantioselectivity trend is observed within the individual solvent families investigated (alkanes, ethers and aromatics). © Canadian Crown Copyright 1997  相似文献   

13.
Tetralin‐1,4‐dione, the stable tautomer of dihydroxynaphthalene, was reduced with catecholborane in the presence of 3,3‐diphenyl‐1‐butyltetrahydro‐3H‐pyrrolo[1,2‐c][1,3,2]oxazaborole as catalyst to give enantiomerically highly enriched 4‐hydroxy‐1‐tetralone (99% ee) in an efficient one‐pot procedure. The R‐enantiomer provided a rapid access to sertraline while the S‐enantiomer was converted into 2‐epicatalponol and catalponol. A more selective enantioselective route to the antithermitic catalponol made use of the planar chiral tricarbonylchromium complex of hydroxytetralone. Its precursor chromium(tricarbonyl)[η6‐(1‐4,4a,8a)‐tetralin‐5,8‐dione] was obtained via direct complexation of 1,4‐dihydroxynaphthalene using chromium(tricarbonyl)‐ tris(ammonia) and boron trifluoride etherate as source of the chromium(tricarbonyl) fragment. Enolate prenylation was best carried out in the presence of a tetraamine ligand. Complete inversion of the stereogenic center bearing the prenyl group of the initially obtained tetralone complex was achieved via enolate formation followed by protonation.  相似文献   

14.
(11bR,11′bR)‐4,4′‐(1,2‐Phenylene)bis[4,5‐dihydro‐3H‐dinaphtho[2,1‐c:1′,2′‐e]phosphepin] [abbreviated as (R)‐BINAPHANE], (3R,3′R,4S,4′S,11bS,11′bS)‐4,4′‐bis(1,1‐dimethylethyl)‐4,4′,5,5′‐tetrahydro‐3,3′‐bi‐3H‐dinaphtho[2,1‐c:1′,2′‐e]phosphepin [(S)‐BINAPINE], (1S,1′S,2R,2′R)‐1,1′‐bis(1,1‐dimethylethyl)‐2,2′‐biphospholane [(S,S,R,R)‐TANGPHOS] and (2R,2′R,5R,5′R)‐1,1′‐(1,2‐phenylene)bis[2,5‐bis(1‐methylethyl)phospholane] [(R,R)‐i‐Pr‐DUPHOS] are C2‐bridged chiral diphosphines that form stable complexes with palladium(II) and platinum(II) containing a five‐membered chelate ring. The Pd(II)‐BINAPHANE catalyst displayed good to excellent enantioselectivities with ee values as high as 99.0% albeit in low yields for the carbonyl‐ene reaction between phenylglyoxal and alkenes. Its Pt(II) counterpart afforded improved yields while retaining satisfactory enantioselectivity. For the carbonyl‐ene reaction between ethyl trifluoropyruvate and alkenes, the Pd(II)‐BINAPHANE catalyst afforded both good yields and extremely high enantioselectivities with ees as high as 99.6%. A comparative study on the Pd(II) catalysts of the four C2‐bridged chiral diphosphines revealed that Pd(II)‐BINAPHANE afforded the best enantioselectivity. The ee values derived from Pd(II)‐BINAPHANE are much higher than those derived from the other three Pd(II) catalysts. A comparison of the catalyst structures shows that the Pd(II)‐BINAPHANE catalyst is the only one that has two bulky (R)‐binaphthyl groups close to the reaction site. Hence it creates a deep chiral space that can efficiently control the reaction behavior in the carbonyl‐ene reactions resulting in excellent enantioselectivity.  相似文献   

15.
Racemic cis‐10‐azatetracyclo[7.2.0.12,6.14,8]tridecan‐11‐one was prepared from homoadamant‐4‐ene by chlorosulfonyl isocyanate addition. The transformation of the β‐lactam to the corresponding β‐amino ester followed by Candida antarctica lipase A‐catalyzed enantioselective (E>>200) N‐acylation with 2,2,2‐trifluoroethyl butanoate afforded methyl (1R,4R,5S,8S)‐5‐aminotricyclo[4.3.1.13,8]undecane‐4‐carboxylate and the (1S,4S,5R,8R)‐butanamide with>99% ee at 50% conversion. Alternatively, transformation of the β‐lactam to the corresponding N‐hydroxymethyl‐β‐lactam and the following Pseudomonas cepacia (currently Burkholderia cepacia) lipase‐catalyzed enantioseletive O‐acylation provided the (1S,4S,6R,9R)‐alcohol (ee=87%) and the corresponding (1R,4R,6S,9S)‐butanoate (ee>99%). In the latter method, competition for the enzyme between the (1R,4R,6S,9S)‐butanoate, 2,2,2‐trifluoroethyl butanoate and the hydrolysis product, butanoic acid, tended to stop the reaction at about 45% conversion and finally gave racemization in the (1S,4S,6R,9R)‐alcohol with time.  相似文献   

16.
The synthesis of the C2‐symmetrical (2R,5R)‐ and (2S,5S)‐2,5‐bis[(S)‐1‐(dibenzylaminoalkyl)]‐1,4‐dioxanes 1 or 2 in enantiopure form is reported. Compounds 1 and 2 were obtained by a completely selective and unusual cyclodimerization of chiral (2R,1′S)‐ or (2S,1′S)‐2‐(1‐aminoalkyl)epoxides 3 or 4 promoted by a mixture of diisopropylamine and boron trifluoride⋅diethyl etherate complex. The structure of the obtained dioxane was established by single‐crystal X‐ray diffraction analysis. A mechanism has been proposed to explain this transformation.  相似文献   

17.
A deeper understanding of the >99 % S‐selective reduction of both isomers of citral catalyzed by NCR ene reductase was achieved by active‐site mutational studies and docking simulation. Though structurally similar, the E/Z isomers of citral showed a significantly varying selectivity response to introduced mutations. Although it was possible to invert (E)‐citral reduction enantioselectivity to ee 46 % (R) by introducing mutation W66A, for (Z)‐citral it remained ≥88 % (S) for all single‐residue variants. Residue 66 seems to act as a lever for opposite binding modes. This was underlined by a W66A‐based double‐mutant library that enhanced the (E)‐citral derived enantioselectivity to 63 % (R) and significantly lowered the S selectivity for (Z)‐citral to 44 % (S). Formation of (R)‐citronellal from an (E/Z)‐citral mixture is a desire in industrial (?)‐menthol synthesis. Our findings pave the way for a rational enzyme engineering solution.  相似文献   

18.
Various ω‐transaminases were tested for the synthesis of enantiomerically pure amines from the corresponding ketones employing D ‐ or L ‐alanine as amino donor and lactate dehydrogenase to remove the side‐product pyruvate to shift the unfavourable reaction equilibrium to the product side. Both enantiomers, (R)‐ and (S)‐amines, could be prepared with up to 99% ee and >99% conversions within 24 h at 50 mM substrate concentration. The activity and stereoselectivity of the amination reaction depended on the ω‐transaminase and substrate employed; furthermore the co‐solvent significantly influenced both the stereoselectivity and activity of the transaminases. Best results were obtained by employing ATA‐117 to obtain the (R)‐enantiomer and ATA‐113 or ATA‐103 to access the (S)‐enantiomer with 15% v v−1 DMSO.  相似文献   

19.
Burkholderia cepacia lipase (BCL) shows high enantioselectivity toward chiral primary alcohols, but this enantioselectivity is often unpredictable, especially for substrates that contain an oxygen at the stereocenter. For example, BCL resolves β‐substituted‐γ‐acetyloxymethyl‐γ‐butyrolactones (acetates of a chiral primary alcohol) by hydrolysis of the acetate, but the enantioselectivity varies with the nature and orientation of the β‐alkyl substituent. BCL favors the (R)‐primary alcohol when the β‐alkyl substituent is hydrogen (E=30) or trans methyl (E=38), but the (S)‐primary alcohol when it is cis methyl (E=145). To rationalize this unusual selectivity, we used a combination of experiments to show the importance of polar interactions and modeling to reveal differences in orientations of the enantiomers. Removal of either the lactone carbonyl in the substrate or the polar side chains in the enzyme by using a related enzyme without these side chains decreased the enantioselectivity at least four‐fold. Modeling revealed that the slow enantiomers do not bind by exchanging the location of two substituents relative to the fast enantiomer. Instead, three substituents remain in the same region, but the fourth substituent, hydrogen, inverts to a new location, like an umbrella in a strong wind. In this orientation the favored stereoisomers have similar shapes, thus accounting for the unusual stereoselectivity. The ratio of catalytically productive orientations for the fast vs. slow enantiomers in a molecular dynamic simulation correlated (R2=0.82) with the degree of enantioselectivity including the case where the enantioselectivity reversed. Weighting this ratio by the ratio of H‐bonds in the polar interaction to account for different binding strengths improved the correlation with the measured enantioselectivity to R2=0.97. The modeling identifies key interactions responsible for high enantioselectivity in this class of substrates.

  相似文献   


20.
A catalytic complex made from [Ir(COD)Cl]2 [di‐μ‐chloro‐bis(1,5‐cyclooctadiene)diiridium(I)] precursor and (S,S)‐f‐Binaphane ((R,R)‐1,1′‐bis{(R)‐4,5‐dihydro‐3H‐dinaphtho[1,2‐c:2′,1′‐e]phosphepino}ferrocene) ligand effectively catalyzed the enantioselective hydrogenation of cyclic imines with high reactivity and good enantioselectivity.  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