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.
The phenylalanine aminomutase (PAM) from Taxus chinensis catalyses the conversion of α‐phenylalanine to β‐phenylalanine, an important step in the biosynthesis of the N‐benzoyl phenylisoserinoyl side‐chain of the anticancer drug taxol. Mechanistic studies on PAM have suggested that (E)‐cinnamic acid is an intermediate in the mutase reaction and that it can be released from the enzyme's active site. Here we describe a novel synthetic strategy that is based on the finding that ring‐substituted (E)‐cinnamic acids can serve as a substrate in PAM‐catalysed ammonia addition reactions for the biocatalytic production of several important β‐amino acids. The enzyme has a broad substrate range and a high enantioselectivity with cinnamic acid derivatives; this allows the synthesis of several non‐natural aromatic α‐ and β‐amino acids in excellent enantiomeric excess (ee >99 %). The internal 5‐methylene‐3,5‐dihydroimidazol‐4‐one (MIO) cofactor is essential for the PAM‐catalysed amination reactions. The regioselectivity of amination reactions was influenced by the nature of the ring substituent.  相似文献   

2.
The asymmetric bromohydroxylation of 2‐aryl‐2‐propen‐1‐ols catalyzed by quinine‐derived bifunctional catalyst has been developed. The regioselectivity was controlled by employing a boronate ester as tether which was formed in situ and enantioselectivity was introduced by taking advantage of a quinine‐derived bifunctional catalyst which activated the boronate ester and N‐bromosuccinimide (NBS) at the same time. Chiral bromohydrin, which is a useful feedstock in organic synthesis, was produced in moderate to excellent enantioselectivity in a two‐step reaction sequence.  相似文献   

3.
Catalytic asymmetric conjugate arylation of racemic 6‐substituted cyclohexenones with arylboronic acids was catalyzed by 3 mol % of chiral amidophosphane‐[RhCl(C2H4)]2 in a 10:1 mixture of 1,4‐dioxane and water at 70 °C to afford a nearly 1:1 mixture of trans‐ and cis‐5‐aryl‐2‐substituted cyclohexanones in high enantioselectivity, which was subsequently epimerized with sodium ethoxide in ethanol to give thermodynamically stable trans‐5‐aryl‐2‐substituted cyclohexanones with 99–97 % ee in high two‐step yields.  相似文献   

4.
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   

5.
The bifunctional quinine‐derived thiourea catalyst 14 was found to catalyze the direct amination of unprotected 3‐aryl and aliphatic substituted oxindoles with di‐tert‐butyl azodicarboxylate (DBAD) to construct a tetrasubstituted stereogenic carbon center at the C‐3 position of oxindoles in good to excellent yield and enantioselectivity.  相似文献   

6.
A more practical and efficient catalytic asymmetric chlorolactonization of styrene‐type carboxylic acids with 1,3‐dichloro‐5,5‐dimethylhydantoin (DCDMH) using C3‐symmetric cinchonine‐squaramide (CSCS) as organocatalyst has been developed. A series of chiral chloro‐substituted isochroman‐1‐ones was obtained in excellent yields (up to 95%) and enantioselectivities (up to 99% ee), whwereby the results for chloro‐substituted isochroman‐1‐ones are the best ever achieved. The catalyst can be recovered and reused for six cycles. Moreover, the chlorolactonization product 3b was further transformed to optically active bicyclic isochroman‐1‐one derivatives in high yield without losing the enantioselectivity. Furthermore, compounds 3e and 2n proved to be highly potent inhibitors of the HIV‐1 in TZM‐bl cells.

  相似文献   


7.
The enantioselectivity displayed by the lipase from Pseudomonas cepacia towards a wide range of prochiral 2‐substituted‐propane‐1,3‐diamines was studied by means of molecular dynamics simulations (MDS). In all cases the enzyme allows the recovery of the corresponding amino carbamates of R configuration. However, the enantioselectivity is only synthetically useful if no ortho substituent is present and the aromatic ring is directly bonded to the 2‐carbon of the 1,3‐diamine core. Analysis of the MDS trajectories revealed that the homologation of 2‐aryl substituents by means of a methylene group lowers enantioselectivity by alleviating the conformational tension of the slow‐reacting orientations due to unfavourable intramolecular contacts between the ortho carbons of the aryl group and the nucleophilic nitrogen, as well as between the chiral carbon and the oxyanion. Additionally, the relative solvent accessible surfaces of the atoms of the aryl ring nicely correlate with the effect of the location of the substituent on enantioselectivity.  相似文献   

8.
Baeyer–Villiger monooxygenase (BVMO)‐mediated regiodivergent conversions of asymmetric ketones can lead to the formation of “normal” or “abnormal” lactones. In a previous study, we were able to change the regioselectivity of a BVMO by mutation of the active‐site residues to smaller amino acids, which thus created more space. In this study, we demonstrate that this method can also be used for other BVMO/substrate combinations. We investigated the regioselectivity of 2‐oxo‐Δ3‐4,5,5‐trimethylcyclopentenylacetyl‐CoA monooxygenase from Pseudomonas putida (OTEMO) for cis‐bicyclo[3.2.0]hept‐2‐en‐6‐one ( 1 ) and trans‐dihydrocarvone ( 2 ), and we were able to switch the regioselectivity of this enzyme for one of the substrate enantiomers. The OTEMO wild‐type enzyme converted (?)‐ 1 into an equal (50:50) mixture of the normal and abnormal products. The F255A/F443V variant produced 90 % of the normal product, whereas the W501V variant formed up to 98 % of the abnormal product. OTEMO F255A exclusively produced the normal lactone from (+)‐ 2 , whereas the wild‐type enzyme was selective for the production of the abnormal product. The positions of these amino acids were equivalent to those mutated in the cyclohexanone monooxygenases from Arthrobacter sp. and Acinetobacter sp. (CHMOArthro and CHMOAcineto) to switch their regioselectivity towards (+)‐ 2 , which suggests that there are hot spots in the active site of BVMOs that can be targeted with the aim to change the regioselectivity.  相似文献   

9.
A series of 1‐alkylbicyclo[3.3.0]octane‐2,8‐diones was transformed by enzymatic retroClaisen reaction using recombinant 6‐oxocamphor hydrolase (OCH) overexpressed in Escherichia coli, to yield optically active 2,3‐substituted cyclopentanones with enantiomeric excesses of up to >95 %. Whilst the parent substrate, bicyclo[3.3.0]octane‐2,8‐dione 12 , was transformed only very slowly, derivatives 13, 14, 15, 16 and 30 with alkyl chains of varying length in the 1‐position were converted rapidly to optically active products with typically 82 % de and up to >95 % enantiomeric excess. The results confirm the apparent requirement of OCH for non‐enolisable diketone substrates, and offer a potential route to decorated cyclopentanone derivatives of multiple chiral centres. Computer modelling of 1‐methylbicyclo[3.3.0]octane‐2,8‐dione into the active site of OCH suggested that the bicyclic [3.3.0] series substrates were accommodated in the active site in similar orientation with the natural enzyme substrate, 6‐oxocamphor, and would thus yield the (2S,3S)‐product series.  相似文献   

10.
A facile chemo‐enzymatic process has been developed for producing stereoisomers of 4‐substituted 2‐hydroxy‐4‐butyrolactones with good to excellent enantioselectivity. This process involves an easy separation of the diastereoisomers by column chromatography and efficient enzymatic resolution by whole cells of Escherichia coli JM109 expressing Fusarium proliferatum lactonase gene. This biocatalyst shows strong tolerance towards different substrate structures and at least three out four possible isomers could be obtained in excellent enantiomeric purity. Different substrate concentrations (10 mM–200 mM) were examined, giving a substrate to catalyst ratio of up to 26:1. This general and efficient enzymatic process provides access to stereoisomers of 4‐substituted 2‐hydroxy‐4‐butyrolactones readily and cost‐effectively. The stereochemical assignments were conducted systematically based on NMR, X‐ray diffraction and circular dichroism, leading to further understanding of the enzyme’s stereoselectivity.  相似文献   

11.
A series of bioisosteric N1‐ and N2‐substituted 5‐(piperidin‐4‐yl)‐3‐hydroxypyrazole analogues of the partial GABAAR agonists 4‐PIOL and 4‐PHP have been designed, synthesized, and characterized pharmacologically. The unsubstituted 3‐hydroxypyrazole analogue of 4‐PIOL ( 2 a ; IC50~300 μM ) is a weak antagonist at the α1β2γ2 GABAAR, whereas substituting the N1‐ or N2‐position with alkyl or aryl substituents resulted in antagonists with binding affinities in the high nanomolar to low micromolar range at native rat GABAARs. Docking studies using a α1β2γ2 GABAAR homology model along with the obtained SAR indicate that the N1‐substituted analogues of 4‐PIOL and 4‐PHP, 2 a – k , and previously reported 3‐substituted 4‐PHP analogues share a common binding mode to the orthosteric binding site in the receptor. Interestingly, the core scaffold of the N2‐substituted analogues of 4‐PIOL and 4‐PHP, 3 b – k , are suggested to flip 180° thereby adapting to the binding pocket and addressing a cavity situated above the core scaffold.  相似文献   

12.
A series of new substituted 7‐phenyl‐3H‐pyrrolo[3,2‐f]quinolin‐9‐ones were synthesized and evaluated for their antiproliferative activity. The most active derivatives showed high selectivity against human leukemia cell lines and potently inhibited their growth, with GI50 values in the nanomolar range. The active compounds strongly blocked tubulin assembly and colchicine binding to tubulin. Their activities were equal to or greater than that of the reference compound combretastatin A‐4. Flow cytometry studies showed that the two most active compounds arrested Jurkat cells in the G2/M cell‐cycle phase in a concentration‐dependent manner. This effect was associated with apoptosis, mitochondrial depolarization, generation of reactive oxygen species, activation of caspase‐3, and cleavage of the enzyme poly(ADP‐ribose) polymerase.  相似文献   

13.
The emergence and spread of antibiotic‐resistant pathogens is a global public health problem. Metallo‐β‐lactamases (MβLs) such as New Delhi MβL‐1 (NDM‐1) are principle contributors to the emergence of resistance because of their ability to hydrolyze almost all known β‐lactam antibiotics including penicillins, cephalosporins, and carbapenems. A clinical inhibitor of MBLs has not yet been found. In this study we developed eighteen new diaryl‐substituted azolylthioacetamides and found all of them to be inhibitors of the MβL L1 from Stenotrophomonas maltophilia (Ki<2 μM ), thirteen to be mixed inhibitors of NDM‐1 (Ki<7 μM ), and four to be broad‐spectrum inhibitors of all four tested MβLs CcrA from Bacteroides fragilis, NDM‐1 and ImiS from Aeromonas veronii, and L1 (Ki<52 μM ), which are representative of the B1a, B1b, B2, and B3 subclasses, respectively. Docking studies revealed that the azolylthioacetamides, which have the broadest inhibitory activity, coordinate to the ZnII ion(s) preferentially via the triazole moiety, while other moieties interact mostly with the conserved active site residues Lys224 (CcrA, NDM‐1, and ImiS) or Ser221 (L1).  相似文献   

14.
The first catalytic asymmetric construction of the cyclic enaminone‐based 3‐substituted 3‐amino‐2‐oxindole scaffold with potential bioactivity has been developed via chiral phosphoric acid‐catalyzed enantioselective addition reactions of cyclic enaminones to isatin‐derived imines, which afforded a series of cyclic enaminone‐based 3‐substituted 3‐amino‐2‐oxindoles in high yields and excellent enantioselectivities (up to 99% yield, 97% ee). The investigation of the reaction mechanism suggested that it was facilitated by a dual hydrogen‐bonding activation mode between the two substrates and the chiral phosphoric acid. Besides, this method could be utilized for a large‐scale synthesis with maintained enantioselectivity. This approach will not only offer a useful method for enantioselective construction of the cyclic enaminone‐based 3‐substituted 3‐amino‐2‐oxindole scaffold, but also enrich the research on catalytic asymmetric addition reactions of isatin‐derived imines by using electron‐rich olefins as nucleophiles. More importantly, a preliminary evaluation on the cytotoxicity of some selected products revealed that two of the enantio‐pure compounds exhibited moderate to strong cytotoxicity to A549, 786‐0, ECA109 and BT474 cancer cell lines.

  相似文献   


15.
Biocatalytic reduction of α‐ or β‐alkyl‐β‐arylnitroalkenes provides a convenient and efficient method to prepare chiral substituted nitroalkanes. Pentaerythritol tetranitrate reductase (PETN reductase) from Enterobacter cloacae st. PB2 catalyses the reduction of nitroolefins such as 1‐nitrocyclohexene ( 1 ) with steady state and rapid reaction kinetics comparable to other old yellow enzyme homologues. Furthermore, it reduces 2‐aryl‐1‐nitropropenes ( 4a–d ) to their equivalent (S)‐nitropropanes 9a–d . The enzyme shows a preference for the (Z)‐isomer of substrates 4a–d , providing almost pure enantiomeric products 9a–d (ees up to>99%) in quantitative yield, whereas the respective (E)‐isomers are reduced with lower enantioselectivity (63–89% ee) and lower product yields. 1‐Aryl‐2‐nitropropenes ( 5a , b ) are also reduced efficiently, but the products (R)‐ 10 have lower optical purities. The structure of the enzyme complex with 1‐nitrocyclohexene ( 1 ) was determined by X‐ray crystallography, revealing two substrate‐binding modes, with only one compatible with hydride transfer. Models of nitropropenes 4 and 5 in the active site of PETN reductase predicted that the enantioselectivity of the reaction was dependent on the orientation of binding of the (E)‐ and (Z)‐substrates. This work provides a structural basis for understanding the mechanism of asymmetric bioreduction of nitroalkenes by PETN reductase.  相似文献   

16.
An organocatalytic Michael addition of protected 2‐amino‐1‐nitroethanes to α,β‐unsaturated aldehydes followed by treatment with TFA afforded 4‐substituted 3‐nitro‐1,2,3,4‐tetrahydropyridines with good diastereoselectivity and excellent enantioselectivity. Good yields were observed in the case of β‐aryl‐substituted α,β‐unsaturated aldehydes as the substrates, while moderate yields were obtained when β‐alkyl‐substituted α,β‐unsaturated aldehydes were used.  相似文献   

17.
APOBEC3G (A3G) is a single‐stranded DNA cytosine deaminase that functions in innate immunity against retroviruses and retrotransposons. Although A3G can potently restrict Vif‐deficient HIV‐1 replication by catalyzing excessive levels of G→A hypermutation, sublethal levels of A3G‐catalyzed mutation may contribute to the high level of HIV‐1 fitness and its incurable prognosis. To chemically modulate A3G catalytic activity with the goal of decreasing the HIV‐1 genomic mutation rate, we synthesized and biochemically evaluated a class of 4‐amino‐1,2,4‐triazole‐3‐thiol small‐molecule inhibitors identified by high‐throughput screening. This class of compounds exhibits low‐micromolar (3.9–8.2 μM ) inhibitory potency and remarkable specificity for A3G versus the related cytosine deaminase, APOBEC3A. Chemical modification of inhibitors, A3G mutational screening, and thiol reactivity studies implicate C321, a residue proximal to the active site, as the critical A3G target for this class of molecules.  相似文献   

18.
Blocking the 2‐C‐methyl‐d ‐erythrithol‐4‐phosphate pathway for isoprenoid biosynthesis offers new ways to inhibit the growth of Plasmodium spp. Fosmidomycin [(3‐(N‐hydroxyformamido)propyl)phosphonic acid, 1 ] and its acetyl homologue FR‐900098 [(3‐(N‐hydroxyacetamido)propyl)phosphonic acid, 2 ] potently inhibit 1‐deoxy‐d ‐xylulose‐5‐phosphate reductoisomerase (Dxr), a key enzyme in this biosynthetic pathway. Arylpropyl substituents were introduced at the β‐position of the hydroxamate analogue of 2 to study changes in lipophilicity, as well as electronic and steric properties. The potency of several new compounds on the P. falciparum enzyme approaches that of 1 and 2 . Activities against the enzyme and parasite correlate well, supporting the mode of action. Seven X‐ray structures show that all of the new arylpropyl substituents displace a key tryptophan residue of the active‐site flap, which had made favorable interactions with 1 and 2 . Plasticity of the flap allows substituents to be accommodated in many ways; in most cases, the flap is largely disordered. Compounds can be separated into two classes based on whether the substituent on the aromatic ring is at the meta or para position. Generally, meta‐substituted compounds are better inhibitors, and in both classes, smaller size is linked to better potency.  相似文献   

19.
Engineered enzyme variants of potato epoxide hydrolase (StEH1) display varying degrees of enrichment of (2R)‐3‐phenylpropane‐1,2‐diol from racemic benzyloxirane. Curiously, the observed increase in the enantiomeric excess of the (R)‐diol is not only a consequence of changes in enantioselectivity for the preferred epoxide enantiomer, but also to changes in the regioselectivity of the epoxide ring opening of (S)‐benzyloxirane. In order to probe the structural origin of these differences in substrate selectivity and catalytic regiopreference, we solved the crystal structures for the evolved StEH1 variants. We used these structures as a starting point for molecular docking studies of the epoxide enantiomers into the respective active sites. Interestingly, despite the simplicity of our docking analysis, the apparent preferred binding modes appear to rationalize the experimentally determined regioselectivities. The analysis also identifies an active site residue (F33) as a potentially important interaction partner, a role that could explain the high conservation of this residue during evolution. Overall, our experimental, structural, and computational studies provide snapshots into the evolution of enantioconvergence in StEH1‐catalyzed epoxide hydrolysis.  相似文献   

20.
The catalytic asymmetric Claisen rearrangement of 2‐alkoxycarbonyl‐substituted allyl vinyl ethers that contain two stereogenic double bonds is described. A combination of the highly Lewis acidic [Cu{(S,S)‐tert‐Bu‐box}](H2O)2(SbF6)2 complex and molecular sieves served as catalyst and afforded the Claisen rearrangement products, substituted and functionalized α‐keto esters, in high yield with a remarkable diastereo‐ and enantioselectivity. The influence of ligand structure, counterion and allyl vinyl ether double bond configuration on the stereoselectivity of the rearrangement was briefly investigated. We propose an explanation for the rate accelerating effect of the Lewis acid as well as a stereochemical model which serve to explain and predict the stereochemical course of the copper bis(oxazoline) catalyzed Claisen rearrangement.  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