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 451 毫秒
1.
We previously reported on enantioselective aldol reactions of acetone and some aldehydes catalyzed by chiral Zn2+ complexes of L ‐prolyl‐pendant [12]aneN4 (L ‐ZnL1) and L ‐valyl‐pendant [12]aneN4 (L ‐ZnL2) in aqueous solution. Here, we report on the one‐pot chemoenzymatic synthesis of chiral 1,3‐diols in an aqueous solvent system at room temperature by a combination of enantioselective aldol reactions catalyzed by Zn2+ complexes of L ‐ and D ‐phenylalanyl‐pendant [12]aneN4 (L ‐ZnL3 and D ‐ZnL3) and the successive enantioselective reduction of the aldol products using oxidoreductases with the regeneration of the NADH (reduced form of nicotinamine adenine dinucleotide) cofactor. The findings indicate that all four stereoisomers of 1,3‐diols can be produced by appropriate selection of a chiral Zn2+‐complex and an oxidoreductase commercially available from the “Chiralscreen OH” kit.  相似文献   

2.
The colorimetric detection of anionic species has been studied for α‐amino acid‐conjugated poly(phenylacetylene)s, which were prepared by the polymerization of the ethyl esters of N‐(4‐ethynylphenylsulfonyl)‐L ‐alanine, L ‐isoleucine, L ‐valine, L ‐phenylalanine, L ‐aspartic acid, and L ‐glutamic acid using Rh+(2,5‐norbornadiene)[(η6‐C6H5)B?(C6H5)3] as the catalyst in CHCl3. The one‐handed helical conformations of all the sulfonamide‐functionalized polymers were characterized by Cotton effects in the circular dichroism spectra. The addition of anions with a relatively high basicity, such as tetra‐n‐butylammonium acetate and fluoride, induced drastic changes in both the optical and chiroptical properties. On the other hand, anions with a relatively low basicity, such as tetra‐n‐butylammonium nitrate, azide, and bromide, had essentially no effects on the helical conformation of all the sulfonamide‐functionalized polymers. The anion signaling property of the sulfonamide‐functionalized polymers possessing α‐amino acid moieties was significantly affected by the installed residual amino acid structures. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 1683–1689, 2010  相似文献   

3.
Three different characterization methods—13C NMR spectroscopy, a terminal terpolymerization model, and a probability analysis based on the Poisson distribution—were used to determine the microstructure of random terpolymers. The methods were used to determine the amino acid sequence distribution of random terpolymers prepared from the polymerization of N‐carboxyanhydrides that contained L ‐leucine, β‐benzyl‐L ‐aspartate, and L ‐valine. Poly(L ‐leucine‐L ‐aspartic acid‐L ‐valine) [poly(LDV)] was designed as a target specific substrate for the α4β1 integrin that recognizes the tripeptide sequence leucine‐aspartic acid‐valine (LDV). The presence of the tripeptide sequence LDV within the polymer was determined to be eight LDV triad sequences on average in terpolymers of approximately 100 kDa. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 4328–4337, 2006  相似文献   

4.
A new hybrid amphiphilic system between a polyacrylic acid (PAA) synthetic segment, and a hydrophobic β‐sheet forming peptide segment, poly(L ‐valine) (PLVAL) was synthesized using a combination of Atom Transfer Radical Polymerization, Click Chemistry, and Nickel catalyzed ring opening of L ‐valine N‐carboxyanhydride. This is the first reported use of Click Chemistry as an intermediary step for the ω‐amino functionalization of polymers to obtain macroinitiators that are free from deactivating or interfering molecules to be used in subsequent Ni‐catalyzed ring opening reaction. The efficiency of the end‐group functionalization in the macroinitiator is about 90%. Three different PAA‐b‐PLVAL hybrid copolymers with molecular weight range of 8000–15,000 and Mw/Mn <1.3 had been prepared by varying the monomer to macroinitiator ratio. In addition, the highest achievable molecular weight in the copolymerization was found to be limited by the solubility of the growing chains. This combined synthetic approach can potentially be extended for the synthesis of a multitude of other peptide hybrid systems, and hence will be of interest in the preparation of peptide hybrid systems. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 2646–2656, 2007  相似文献   

5.
Eight water‐soluble 1,1′‐dimethylvanadocene amino acid complexes have been prepared via the reaction of (MeCp)2VCl2 ( 2 ) with one equivalent of amino acid (aa) in water affording [(MeCp)2V( aa )]Cl, where aa is glycine ( 3 ), L ‐alanine ( 4 ), L ‐valine ( 5 ), L ‐leucine ( 6 ), L ‐isoleucine ( 7 ), L ‐phenylalanine ( 8 ), L ‐histidine ( 9 ) and L ‐tryptophane ( 10 ). All prepared complexes have been characterized by EPR, IR and Raman spectroscopy, elemental analysis and mass spectrometry. Molecular structures of [(MeCp)2V(ala)]BPh4·CH3OH ( 11 ), [(MeCp)2V(leu)]PF6 ( 12 ) and [(MeCp)2V(ile)]PF6 ( 13 ) were determined by X‐ray diffraction analysis. Cytotoxic properties of complexes 2–10 were investigated toward Escherichia coli B and compared with analogical unsubstituted vanadocene compounds ( 1, 14–21 ). The results showed that 1,1′‐dimethylvanadocene amino acid complexes have identical or slightly higher antiproliferative activity then their unsubstituted analogs. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

6.
The reaction of 2,4‐pentanedione ( 1 ) with (R)‐(—)‐2‐phenylglycine methyl ester ( 2 ), (R)‐(—)‐2‐phenylglycinol ( 3 ) and the proteinogenic amino acids (2S,3R)‐(—)‐2‐amino‐3‐hydroxybutyric acid (L ‐threonine) ( 4 ) and (R)‐(—)‐2‐amino‐3‐mercaptopropionic acid (L ‐cysteine) ( 5 ) methyl esters was investigated. The corresponding enamines 6 , 7 and 8 were isolated and characterized spectroscopically whereas 9 , which is unstable, was transformed in situ into 13 . Treatment of 7 , 8 and 9 with boron trifluoride etherate afforded the new [1,4]oxazepines 10 , 11 and [1,4]thiazepine ( 12 ) as their BF3O? salts. The structures of the enamines and their corresponding seven‐membered heterocycles were assessed by 1D and 2D NMR spectroscopy. Variable‐temperature experiments revealed different molecular mobility behavior among these heterocycles. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

7.
To study the bidentate coordination effect on the polycondensation of L ‐valinates between metal triflates as a Lewis acid and methoxy groups, we carried out the polycondensation of 2‐methoxy‐4‐nitrophenyl L ‐valinate ( 1a ) and 2‐methoxyphenyl L ‐valinate ( 1b ) in the presence of the various kinds of rare‐earth triflates in DMF solution at room temperature. The polymerizations of 1a did not proceed without any metal triflates. In the presence of 5 mol% triflates, especially Sc(OTf)3, the polymerization proceeded effectively. After the reaction mixture was poured into water, the product was collected, which was recognized as poly(L ‐valine)s by FTIR spectrum and GPC measurement. The yield of the product from the polymerization of 1a with Sc(OTf)3 was higher than that from the polymerization of 4‐nitrophenyl L ‐valinate ( 1c ) with Sc(OTf)3. This result indicates that the polymerization of 1a was promoted to introduce the methoxy group on the o‐position of the phenyl ring at the ester group with the aim of the bidentate coordination effect between metal triflates and L ‐valinate. As a control experiment, we carried out the polycondensation of 1b in the presence of 5 mol% metal triflates; however, any polymerization did not proceeded. That reason is from the lower activity of activated L ‐valinate ( 1b ). © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 2864–2868, 2008  相似文献   

8.
Dipeptides with two hydrophobic side chains have proved to be an exceptional source of microporous organic materials, but since previous structures were limited to the incorporation of only proteinogenic residues, their full potential as adsorbents has remained unexplored. Single‐crystal XRD data for ten new compounds with non‐proteinogenic L ‐2‐aminobutanoic acid and/or L ‐2‐amino‐pentanoic acid are presented. The gas‐phase accessibility of their crystal pores, with cross‐sections of 2.3 to 5.1 Å, was monitored by CO2 and CH4 adsorption isotherms. Included CO2 was also detected spectroscopically by 2D MAS NMR. An extensive conformational analysis reveals that the use of linear rather than branched side chains (such as L ‐valine and L ‐isoleucine) affords peptides with a greater degree of conformational freedom and yields more‐flexible channel surfaces that may easily adapt to a series of potential guest molecules.  相似文献   

9.
Novel polyphenylacetylene and polystyrene derivatives carrying L ‐proline moieties at the side chains were synthesized by the rhodium‐catalyzed and radical polymerizations of the corresponding monomers. The polyphenylacetylene derivatives showed Cotton effects at the absorption region of the main chain, indicating that the polymers adopt helical conformations with predominantly one‐handed screw sense. The polymers catalyzed the asymmetric aldol reactions of acetone with aromatic aldehydes, and cyclohexanone with p‐nitrobenzaldehyde. The enantioselectivities largely depended on the reaction conditions. In the asymmetric aldol reaction of acetone with aromatic aldehydes, the R‐enantiomeric products were predominantly obtained except the cases with the polymer catalyst in CHCl3. The ee of the products became higher as the reaction temperature was decreased. The polymeric catalysts were recoverable from the reaction mixture by filtration, and the recovered ones catalyzed the asymmetric aldol reaction of acetone with p‐nitrobenzaldehyde without decreasing the product yield and ee. The ee was improved using the copolymers of L ‐proline‐based and nonchiral monomers as catalysts. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2011  相似文献   

10.
The bicyclic depsipeptide histone deacetylase (HDAC) inhibitors spiruchostatins A and B, 5′′‐epi‐spiruchostatin B and FK228 were efficiently synthesized in a convergent and unified manner. The synthetic method involved the following crucial steps: i) a Julia–Kocienski olefination of a 1,3‐propanediol‐derived sulfone and a L ‐ or D ‐malic acid‐derived aldehyde to access the most synthetically challenging unit, (3S or 3R,4E)‐3‐hydroxy‐7‐mercaptohept‐4‐enoic acid, present in a D ‐alanine‐ or D ‐valine‐containing segment; ii) a condensation of a D ‐valine‐D ‐cysteine‐ or D ‐allo‐isoleucine‐D ‐cysteine‐containing segment with a D ‐alanine‐ or D ‐valine‐containing segment to directly assemble the corresponding seco‐acids; and iii) a macrocyclization of a seco‐acid using the Shiina method or the Mitsunobu method to construct the requisite 15‐ or 16‐membered macrolactone. The present synthesis has established the C5′′ stereochemistry of spiruchostatin B. In addition, HDAC inhibitory assay and the cell‐growth inhibition analysis of the synthesized depsipeptides determined the order of their potency and revealed some novel aspects of structure–activity relationships. It was also found that unnatural 5′′‐epi‐spiruchostatin B shows extremely high selectivity (ca. 1600‐fold) for class I HDAC1 (IC50=2.4 nM ) over class II HDAC6 (IC50=3900 nM ) with potent cell‐growth‐inhibitory activity at nanomolar levels of IC50 values.  相似文献   

11.
To develop a novel polycondensation method for the preparation of poly (amino acid)s, we screened a transition metal or a rare‐earth triflate as a Lewis acid for the polycondensation of activated amino acid esters in N,N‐dimethylformamide solutions at room temperature. The polymerizations of 4‐nitrophenyl L ‐leucinate ( 1a ) and 4‐nitrophenyl L ‐valinate ( 1b ) scarcely proceeded without any Lewis acid at room temperature. In the presence of 5 mol % metal triflates, especially scandium(III) trifluoromethanesulfonate, the polymerizations of both monomers were promoted effectively. The products, which were collected by the reaction mixture being poured into water, were recognized as poly(L ‐valine)s by Fourier transform infrared spectroscopy, gel permeation chromatography analysis, and 1H NMR spectroscopy. These results showed that a metal triflate as a Lewis acid could coordinate to a carbonyl oxygen of activated L ‐valinate and L ‐leucinate even in a highly polar solvent, such as N,N‐dimethylformamide; therefore, the polymerizations of activated L ‐valinate and L ‐leucinate were promoted. Because steric hindrance derived from the isobutyl group in 1b was less than that of the isopropyl unit in 1a , the effect of the metals was not as sensitive for the polymerization of 1b . © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 543–547, 2007  相似文献   

12.
Amino acid based thioamides, hydroxamic acids, and hydrazides have been evaluated as ligands in the rhodium‐catalyzed asymmetric transfer hydrogenation of ketones in 2‐propanol. Catalysts containing thioamide ligands derived from L ‐valine were found to selectively generate the product with an R configuration (95 % ee), whereas the corresponding L ‐valine‐based hydroxamic acids or hydrazides facilitated the formation of the (S)alcohols (97 and 91 % ee, respectively). The catalytic reduction was examined by performing a structure–activity correlation investigation with differently functionalized or substituted ligands and the results obtained indicate that the major difference between the thioamide and hydroxamic acid based catalysts is the coordination mode of the ligands. Kinetic experiments were performed and the rate constants for the reduction reactions were determined by using rhodium–arene catalysts derived from amino acid thioamide and hydroxamic acid ligands. The data obtained show that the thioamide‐based catalyst systems demonstrate a pseudo‐first‐order dependence on the substrate, whereas pseudo‐zero‐order dependence was observed for the hydroxamic acid containing catalysts. Furthermore, the kinetic experiments revealed that the rate‐limiting steps of the two catalytic systems differ. From the data obtained in the structure–activity correlation investigation and along with the kinetic investigation it was concluded that the enantioswitchable nature of the catalysts studied originates from different ligand coordination, which affects the rate‐limiting step of the catalytic reduction reaction.  相似文献   

13.
Histidine ammonia‐lyase (HAL) was irreversibly inhibited by L ‐cysteine at pH 10.5 under aerobic conditions. The inhibited enzyme, still in its intact conformation, showed an absorption maximum at 338 nm. Upon denaturation, followed by pronase digestion, two main chromophoric products 1 and 2 (Figs. 4 and 5, resp.) could be isolated with absorption maxima at 335 and 332 nm, respectively. As determined by MALDI‐TOF mass spectrometry and 1H‐NMR spectroscopy, in product 1 one of the methylidene H‐atoms of the 3,5‐dihydro‐5‐methylidene‐4H‐imidazol‐4‐one (formerly called 4‐methylideneimidazol‐5‐one; MIO) prosthetic group was substituted by one of the amino groups of L ‐ cystine, while in product 2 the ε‐amino group of L ‐lysine was the analogous substituent. Acid‐catalyzed hydrolysis of product 1 gave compound 3 whose chromophore (λmax 310 nm) was that of 3,5‐dihydro‐5‐(4‐hydroxymethylidene)‐4H‐imidazol‐4‐one, i.e., of a vinylogous acid. These results support our previous proposal that, in the first step, the L ‐cysteine S‐atom attacks the prosthetic electrophile (Scheme 2). The resulting nucleophilic enolate captures O2 to form a peroxide. On the basis of the present results, we postulate that the observed products 1 – 3 arise from a vinylogous thioester 4 , which is formed in the conformationally intact inhibited enzyme by an electrocyclic reaction eliminating H2O2.  相似文献   

14.
Two types of three‐arm and four‐arm, star‐shaped poly(D,L ‐lactic acid‐alt‐glycolic acid)‐b‐poly(L ‐lactic acid) (D,L ‐PLGA50‐b‐PLLA) were successfully synthesized via the sequential ring‐opening polymerization of D,L ‐3‐methylglycolide (MG) and L ‐lactide (L ‐LA) with a multifunctional initiator, such as trimethylolpropane and pentaerythritol, and stannous octoate (SnOct2) as a catalyst. Star‐shaped, hydroxy‐terminated poly(D,L ‐lactic acid‐alt‐glycolic acid) (D,L ‐PLGA50) obtained from the polymerization of MG was used as a macroinitiator to initiate the block polymerization of L ‐LA with the SnOct2 catalyst in bulk at 130 °C. For the polymerization of L ‐LA with the three‐arm, star‐shaped D,L ‐PLGA50 macroinitiator (number‐average molecular weight = 6800) and the SnOct2 catalyst, the molecular weight of the resulting D,L ‐PLGA50‐b‐PLLA polymer linearly increased from 12,600 to 27,400 with the increasing molar ratio (1:1 to 3:1) of L ‐LA to MG, and the molecular weight distribution was rather narrow (weight‐average molecular weight/number‐average molecular weight = 1.09–1.15). The 1H NMR spectrum of the D,L ‐PLGA50‐b‐PLLA block copolymer showed that the molecular weight and unit composition of the block copolymer were controlled by the molar ratio of L ‐LA to the macroinitiator. The 13C NMR spectrum of the block copolymer clearly showed its diblock structures, that is, D,L ‐PLGA50 as the first block and poly(L ‐lactic acid) as the second block. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 40: 409–415, 2002  相似文献   

15.
A series of chiral Schiff bases ( L 1 – L 5 ) with different substituents in the salicylidenyl unit were prepared from condensation of 3‐aryl‐5‐ tert ‐butylsalicylaldehyde derivatives and optically active amino alcohols. Bromination of 3‐phenyl‐5‐ tert ‐butylsalicylaldehyde gave an unexpected product 3‐(4‐bromophenyl)‐5‐bromosalicylaldehyde, from which the corresponding Schiff base ligands L 6 and L 7 , derived from (S)‐valinol and (S)‐ tert ‐leucinol, respectively, were prepared. Ligands L 1 – L 7 were applied to the vanadium‐catalyzed asymmetric oxidation of aryl methyl sulfides. Under the optimal conditions, the oxidation of the thioanisole with H2O2 as oxidant in CH2Cl2 catalyzed by VO(acac)2‐ L 1 – L 7 gives good yields (74–83%) with moderate enantioselectivity (58–77% ee). Ligand L 7 , containing a 4‐bromophenyl group on the 3‐position and a Br atom on the 5‐position of the salicylidenyl moiety, displays an 80–90% ee for vanadium‐catalyzed oxidation of methyl 4‐bromophenyl sulfide and methyl 2‐naphthyl sulfide. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

16.
The NCAs of the following five amino acids were polymerized in bulk at 120 °C without addition of a catalyst or initiator: sarcosine (Sar), L ‐alanine (L ‐Ala), D ,L ‐phenylalanine (D ,L ‐Phe), D ,L ‐leucine (D ,L ‐Leu) and D ,L ‐valine (D,L ‐Val). The virgin reaction products were characterized by viscosity measurements 13C NMR spectroscopy and MALDI‐TOF mass spectrometry. In addition to numerous low molar mass byproducts cyclic polypeptides were formed as the main reaction products in the mass range above 800 Da. Two types of cyclic oligo‐ and polypeptides were detected in all cases with exception of sarcosine NCA, which only yielded one class of cyclic polypeptides. The efficient formation of cyclic oligo‐ and polypeptides explains why high molar mass polymers cannot be obtained by thermal polymerizations of α‐amino acid NCAs. Various polymerization mechanisms were discussed. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 4012–4020, 2008  相似文献   

17.
In this article, we describe the one‐step synthesis of polyesters having pendent hydroxyl groups by Lewis acid‐catalyzed, regioselective, dehydration polycondensations of diols (glycerol and sorbitol) and dicarboxylic acids [tartaric acid (TA) and malic acid (MA)] containing pendent hydroxyl groups, using low temperature polycondensation technique. Direct polycondensations of TA or MA and 1,9‐nonanediol catalyzed by scandium trifluoromethanesulfonate [Sc(OTf)3] successfully yielded linear polyesters having hydroxyl functionality (Mn = ca. 1.0 × 104). To demonstrate the reactivity of the pendent hydroxyl group, a glycosidation was performed. Poly(nonamethylene L ‐malate) showed significant higher biodegradability, compared with poly(nonamethylene L ‐tartrate) or poly(nonamethylene succinate). Stable poly(nonamethylene L ‐tartrate) emulsion could be prepared using poly(vinyl alcohol) as the surfactant, although emulsions consisting of poly(nonamethylene succinate) were unstable and phase‐separated within a few days. Furthermore, direct polycondensations of TA and diethylene glycol (DEG) or triethylene glycol (TEG) successfully produced water‐soluble polyesters having hydroxyl groups. This new polycondensation system may be extremely effective not only for advanced material design using functional monomers but also for effective utilization of biomass resources as chemical substances. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 5747–5759, 2009  相似文献   

18.
The Knoevenagel reactions of malononitrile with acetophenone or 4‐substituted acetophenons were carried to give the corresponding 2‐(1‐aryle thylidene)malononitriles, which was further cyclized with sulfur using NaHCO3 as catalysts to generate 2‐amino‐5‐arylthiophene‐3‐carbonitrile 2 . The intermediate enamines 3 were prepared by refluxing of 2 with 5‐substituted‐1,3‐cyclohexanedione using p‐toluenesulfonic acid as catalyst. The title compounds 4‐amino‐3‐aryl ‐7‐substituted‐7,8‐dihydrothieno[2,3‐b]quinolin‐5(6H)‐one were synthesized by cyclization of 3 in the presence of K2CO3 and Cu2Cl2. The structures of all compounds were characterized by elemental analysis, IR, MS, and 1H‐NMR spectra.  相似文献   

19.
1-Morpholino-cyclopenten-2-carboxylic acid anilides react with CF3SCl in molar ratio 1:2 to give products of the displacement of two hydrogen atoms at C-2 and C-5. The analogous reaction with enamines of cyclohexanono- and cycloheptanono- 2-carboxylic acid anilides and benzoylacetanilides proceeds in 1:1 molar ratios affording enamines and keto compounds.  相似文献   

20.
We prepared two new linkers, S‐functionalized adamantane derivatives 2 and 3 , which bind as monolayers on polycrystalline gold. From these surface anchors, both L ‐ and D ‐isomers of alanine can be grown as thin films of α‐helical polypeptides directed from the gold surface by using the appropriate N‐carboxyalanine anhydride. FT‐IR Studies show that these layers are roughly 1000‐Å thick and that, under the same growth conditions, the L ‐polypeptide layers grow at a rate ca. 30% greater than that of the non‐natural D ‐amino acid. X‐Ray photoelectron spectroscopy studies show that, upon equilibration, all three S‐atoms of the sulfide moieties of 2 are bound to the gold surface, and that, on average, three of the four thiols of 3 are chemoadsorbed. The essential role of H2O on the surface of these films as a necessary component in these gas‐phase polymerization reactions is demonstrated.  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