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 half reactions of ω‐aminotransferase (ω‐AT) from Vibrio fluvialis JS17 (ω‐ATVf) were carried out using purified pyridoxal 5′‐phosphate‐enzyme (PLP‐Enz) and pyridoxamine 5′‐phosphate‐enzyme (PMP‐Enz) complexes to investigate the relative activities of substrates. In the reaction generating PMP‐Enz from PLP‐Enz using L ‐alanine as an amine donor, L ‐alanine showed about 70% of the initial reaction rate of (S)‐α‐methylbenzylamine ((S)‐α‐MBA). However, in the subsequent half reaction recycling PLP‐Enz from PMP‐Enz using acetophenone as an amine acceptor, acetophenone showed nearly negligible reactivity compared to pyruvate. These results indicate that the main bottleneck in the asymmetric synthesis of (S)‐α‐MBA lies not in the amination of PLP by alanine, but in the amination of acetophenone by PMP‐Enz, where conformational restraints of the enzyme structure is likely to be the main reason for limiting the amine group transfer from PMP‐Enz to acetophenone. Based upon those half reaction experiments using the two amino acceptors of different activity, it appears that the relative activities of the two amine donors and the two acceptors involved in the ω‐AT reactions can roughly determine the asymmetric synthesis yield of the target chiral amine compound. Predicted conversion yields of several target chiral amines were calculated and compared with the experimental conversion yields. Approximately, a positive linear correlation (Pearson's correlation coefficient = 0.92) was observed between the calculated values and the experimental conversion yields. To overcome the low (S)‐α‐MBA productivity of ω‐ATVf caused by the possible disadvantageous structural constraints for acetophenone, new ω‐ATs showing higher affinity to benzene ring of acetophenone than ω‐ATVf were computationally screened using comparative modeling and protein‐ligand docking. ω‐ATs from Streptomyces avermitilis MA‐4680 (SAV2612) and Agrobacterium tumefaciens str. C58 (Atu4761) were selected, and the two screened ω‐ATs showed higher asymmetric synthesis reaction rate of (S)‐α‐MBA and lower (S)‐α‐MBA degradation reaction rate than ω‐ATVf. To verify the higher conversion yield of the variants of ω‐ATs, the reaction with 50 mM acetophenone and 50 mM alanine was performed with coupling of lactate dehydrogenase and two‐phase reaction system. SAV2612 and Atu4761 showed 70% and 59% enhanced yield in the synthesis of (S)‐α‐MBA compared to that of ω‐ATVf, respectively. Biotechnol. Bioeng. 2011;108: 253–263. © 2010 Wiley Periodicals, Inc.  相似文献   

2.
Four novel chiral amino alcohols were synthesized from D‐(+)‐camphor and utilized as ligands in a Cu(I)‐catalyzed asymmetric Henry reaction. The reactions were carried out under mild conditions with excellent enantioselectivities and moderate yields without the exclusion of air or moisture. The highest enantioselectivity was observed up to 94% enantiomeric excess (ee) with ligand L1 in toluene at room temperature. Chirality 27:761–765, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

3.
Liquid chromatographic enantiomer separation of several N‐benzyloxycarbonyl (CBZ) and Ntert‐butoxycarbonyl (BOC) α‐amino acids and their corresponding ethyl esters was performed on covalently immobilized chiral stationary phases (CSPs) (Chiralpak IA and Chiralpak IB) and coated‐type CSPs (Chiralpak AD and Chiralcel OD) based on polysaccharide derivatives. The solvent versatility of the covalently immobilized CSPs in enantiomer separation of N‐CBZ and BOC‐α‐amino acids and their ester derivatives was shown and the chromatographic parameters of their enantioselectivities and resolution factors were greatly influenced by the nature of the mobile phase. In general, standard mobile phases using 2‐propanol and hexane on Chiralpak IA showed fairly good enantioselectivities for resolution of N‐CBZ and BOC‐α‐amino acids and their esters. However, 50% MTBE/hexane (v/v) for resolution of N‐CBZ‐α‐amino acids ethyl esters and 20% THF/hexane (v/v) for resolution of N‐BOC‐α‐amino acids ethyl esters afforded the greatest enantioselectivities on Chiralpak IA. Also, liquid chromatographic comparisons of the enantiomer resolution of these analytes were made on amylose tris(3,5‐dimethylphenylcarbamate)‐derived CSPs (Chiralpak IA and Chiralpak AD) and cellulose tris(3,5‐dimethylphenylcarbamate)‐derived CSPs (Chiralpak IB and Chiralcel OD). Chiralpak AD and/or Chiralcel OD showed the highest enantioselectivities for resolution of N‐CBZ‐α‐amino acids and esters, while Chiralpak AD or Chiralpak IA showed the highest resolution of N‐BOC‐α‐amino acids and esters. Chirality 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

4.
N‐(tert‐butyloxycarbonyl) or N‐(9‐fluorenylmethoxycarbonyl) dipeptides with C‐terminal (Z)‐α,β‐didehydrophenylalanine (?ZPhe), (Z)‐α,β‐didehydrotyrosine (?ZTyr), (Z)‐α,β‐didehydrotryptophan (?ZTrp), (Z)‐α,β‐didehydromethionine (?ZMet), (Z)‐α,β‐didehydroleucine (?ZLeu), and (Z/E)‐α,β‐didehydroisoleucine (?Z/EIle) were synthesised from their saturated analogues via oxidation of intermediate 2,5‐disubstituted‐oxazol‐5‐(4H)‐ones (also known as azlactones) with pyridinium tribromide followed by opening of the produced unsaturated oxazol‐5‐(4H)‐one derivatives in organic‐aqueous solution with a catalytic amount of trifluoroacetic acid or by a basic hydrolysis. In all cases, a very strong preference for Z isomers of α,β‐didehydro‐α‐amino acid residues was observed except of the ΔIle, which was obtained as the equimolar mixture of Z and E isomers. Reasons for the (Z)‐stereoselectivity and the increased stability of the aromatic α,β‐didehydro‐α‐amino acid residue oxazol‐5‐(4H)‐ones over the corresponding aliphatic ones are also discussed. It is the first use of such a procedure to synthesise peptides with the C‐terminal unsaturated residues and a peptide with 2 consecutive ΔPhe residues. This approach is very effective especially in the synthesis of peptides with aliphatic α,β‐didehydro‐α‐amino acid residues that are difficult to obtain by other methods. It allowed the first synthesis of the ?Met residue. It is also more cost‐effective and less laborious than other synthesis protocols. The dipeptide building blocks obtained were used in the solid‐phase synthesis of model peptides on a polystyrene‐based solid support. Peptides containing aromatic α,β‐didehydro‐α‐amino acid residues were obtained with PyBOP or TBTU as a coupling agent with good yields and purities. In the case of aliphatic α,β‐didehydro‐α‐amino acid residues, a good efficiency was achieved only with DPPA as a coupling agent.  相似文献   

5.
The Pictet–Spengler (PS) cyclizations of β3hTrp derivatives as arylethylamine substrates were performed with L‐α‐amino and D‐α‐amino aldehydes as carbonyl components. During the PS reaction, a new stereogenic center was created, and the mixture of cis/trans 1,3‐disubstituted 1,2,3,4‐tetrahydro‐β‐carbolines was obtained. The ratio of cis/trans diastereomers depends on the stereogenic centre of used amino aldehyde and the size of substituents. It was confirmed by 1H and 2D NMR (ROESY) spectra. The conformations of cyclic products were studied by 2D NMR ROESY spectra. Products of the PS condensation after removal of protecting group(s) can be incorporated into a peptide chain as tryptophan mimetics with the possibility of the β‐turn induction. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
Myung Ho Hyun 《Chirality》2015,27(9):576-588
Crown ether‐based chiral stationary phases (CSPs) have been known to be useful for the resolution of racemic primary amino compounds. In particular, CSPs based on (+)‐(18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid have been reported to be useful for the resolution of secondary amino compounds as well as primary amino compounds. In this article, the process of developing various CSPs based on (+)‐(18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid to improve the chiral recognition efficiency and/or the stability of the CSPs and their applications to the resolution of various primary and nonprimary amino compounds are reviewed. Chirality 27:576588, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

7.
This work reports an efficient Lewis acid catalysed N‐methylation procedure of lipophilic α‐amino acid methyl esters in solution phase. The developed methodology involves the use of the reagent system AlCl3/diazomethane as methylating agent and α‐amino acid methyl esters protected on the amino function with the (9H‐fluoren‐9‐yl)methanesulfonyl (Fms) group. The removal of Fms protecting group is achieved under the same conditions to those used for Fmoc removal. Thus the Fms group can be interchangeable with the Fmoc group in the synthesis of N‐methylated peptides using standard Fmoc‐based strategies. Finally, the absence of racemization during the methylation reaction and the removal of Fms group were demonstrated by synthesising a pair of diastereomeric dipeptides. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
The presence of aminoacylase activities was investigated in a crude extract of Streptomyces ambofaciens ATCC23877. First activities catalyzing the hydrolysis of N‐α or ε‐acetyl‐L‐lysine were identified. Furthermore, the acylation of lysine and different peptides was studied and compared with results obtained with lipase B of Candida antarctica (CALB). Different regioselectivities were demonstrated for the two classes of enzymes. CALB was able to catalyze acylation only on the ε‐position whereas the crude extract from S. ambofaciens possessed the rare ability to catalyze the N‐acylation on the α‐position of the lysine or of the amino‐acid in N‐terminal position of peptides. Two genes, SAM23877_1485 and SAM23877_1734, were identified in the genome of Streptomyces ambofaciens ATCC23877 whose products show similarities with the previously identified aminoacylases from Streptomyces mobaraensis. The proteins encoded by these two genes were responsible for the major aminoacylase hydrolytic activities. Furthermore, we show that the hydrolysis of N‐α‐acetyl‐L‐lysine could be attributed to the product of SAM23877_1734 gene.  相似文献   

9.
In order to synthesize interferon‐γ glycoform having an oligosaccharide at the 97 position by a semi‐synthetic method, interferon‐γ‐polypeptide‐(1–94)‐α‐hydrazide was prepared by the specific Cys‐cyanylation of polypeptide‐(1–94)‐Cys‐His6 expressed from E. coli and subsequent hydrazinolysis in 22% yield (two steps). This polypeptide‐α‐hydrazide was then converted into corresponding polypeptide‐α‐thioester under NaNO2/acid conditions followed by thiolysis in 83% yield. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Tumour necrosis factor‐α (TNF‐ α)is a major contributor to the pathogenesis of insulin resistance associated with obesity and type 2 diabetes. It has been found that endogenous hydrogen sulfide (H2S) contributes to the pathogenesis of diabetes. We have hypothesized that TNF‐α‐induced insulin resistance is involved in endogenous H2S generation. The aim of the present study is to investigate the role of endogenous H2S in TNF‐α‐induced insulin resistance by studying 3T3‐L1 adipocytes. We found that treatment of 3T3‐L1 adipocytes with TNF‐α leads to deficiency in insulin‐stimulated glucose consumption and uptake and increase in endogenous H2S generation. We show that cystathionine γ‐lyase (CSE) is catalysed in 3T3‐L1 adipocytes to generate H2S and that CSE expression and activity are upregulated by TNF‐α treatment. Inhibited CSE by its potent inhibitors significantly attenuates TNF‐α‐induced insulin resistance in 3T3‐L1 adipocytes, whereas H2S treatment of 3T3‐L1 adipocytes impairs insulin‐stimulated glucose consumption and uptake. These data indicate that endogenous CSE/H2S system contributes to TNF‐α‐caused insulin resistance in 3T3‐L1 adipocytes. Our findings suggest that modulation of CSE/H2S system is a potential therapeutic avenue for insulin resistance. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

11.
A variety of applications of 8‐alkynylated nucleosides has prompted the synthesis of new purine analogues. Bromination of unprotected 2‐amino‐2′‐deoxyadenosine with Br2/AcOH/AcONa gives 2‐amino‐8‐bromo‐2′‐deoxyadenosine (87%). The brominated derivative is converted to 8‐alkynylated 2‐amino‐2′‐deoxyadenosines by palladium‐catalyzed Sonogashira cross‐coupling reaction via microwave assistance (81 – 95%). The resulting compounds are further transformed to 8‐alkynylated 2′‐deoxyisoguanosines (52 – 70%). The physical properties of new compounds are investigated.  相似文献   

12.
Readily available L‐tartaric acid, which is a bidentate ligand with two chiral centers forming a seven‐membered chelate ring, was applied to the chiral ligand for the chiral nuclear magnetic resonance (NMR) shift reagent of samarium(III) formed in situ. This simple method does not cause serious signal broadening in the high magnetic field. Enantiomeric 13C and 1H NMR signals and enantiotopic 1H NMR signals of α‐amino acids were successfully resolved at pH 8.0 and the 1:3 molar ratio of Sm(NO3)3:L‐tartaric acid. It is elucidated that the enantiomeric signal resolution is attributed to the anisotropic magnetic environment for the enantiomers induced by the chiral L‐tartarato samarium(III) complex rather than differences in stability of the diastereomeric substrate adducts. The present 13C NMR signal resolution was also effective for the practical simultaneous analysis of plural kinds of DL‐amino acids. Chirality 27:353–357, 2015.© 2015 Wiley Periodicals, Inc.  相似文献   

13.
The compound (18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid was evaluated as a chiral nuclear magnetic resonance (NMR) solvating agent for a series of diamines and bicyclic β‐amino acids. The amine must be protonated for strong association with the crown ether. An advantage of (18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid over many other crown ethers is that it undergoes a neutralization reaction with neutral amines to form the protonated species needed for binding. Twelve primary diamines in neutral and protonated forms were evaluated. Diamines with aryl and aliphatic groups were examined. Some are atropisomers with equivalent amine groups. Others have two nonequivalent amine groups. Association equilibria for these systems are complex, given the potential formation of 2:1, 1:1, and 1:2 crown‐amine complexes and given the various charged species in solution for mixtures of the crown ether with the neutral amine. The crown ether produced enantiomeric differentiation in the 1H NMR spectrum of one or more resonances for every diamine substrate. Also, a series of five bicyclic β‐amino acids were examined and (18‐crown‐6)‐2,3,11,12‐tetracarboxylic acid caused enantiomeric differentiation in the 1H NMR spectrum of three or more resonances of each compound. Chirality 27:708–715, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

14.
15.
Chiral functionalization of 2,4,5,6‐tetrachloro‐1,3‐dicyanobenzene (1) by regioselective nucleophilic substitution of one or two chlorine atoms by optically pure (R)‐(+)‐1‐naphthylethylamine (NEA), or by a glycine unit as a spacer to (R)‐NEA, enables the preparation of brush‐type chiral selectors (2, 3, 9, 13). By the introduction of the 3‐aminopropyltriethoxysilyl (APTES) group, reactive intermediates 4a/b, 5, 10a/b, and 14a/b are obtained ( a/b indicate a mixture of regioisomers with APTES in 6‐ and 2‐position). Binding of these to silica gel afforded four novel chiral stationary phases (CSPs) 6, 7, 15, and 16. HPLC columns containing CSPs with (R)‐NEA directly linked to polysubstituted aromatic ring (6, 7) are not very effective in resolution of most of the 23 racemic analytes, whereas the columns with distant π‐basic subunits (15, 16) exhibited higher resolving efficacy, in particular towards the isopropyl esters of racemic N‐3,5‐dinitrobenzoyl‐α‐amino acids. Effective resolution of test racemates reveals the importance of the presence of the hydrogen bond donor amido group and the distance between the persubstituted benzene ring in 1 and the π‐basic naphthalene ring of (R)‐NEA. Chirality 11:722–730, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

16.
Terpenes are important compounds in plant trophic interactions. A meta‐analysis of GC‐MS data from a diverse range of apple (Malus × domestica) genotypes revealed that apple fruit produces a range of terpene volatiles, with the predominant terpene being the acyclic branched sesquiterpene (E,E)‐α‐farnesene. Four quantitative trait loci (QTLs) for α‐farnesene production in ripe fruit were identified in a segregating ‘Royal Gala’ (RG) × ‘Granny Smith’ (GS) population with one major QTL on linkage group 10 co‐locating with the MdAFS1 (α‐farnesene synthase‐1) gene. Three of the four QTLs were derived from the GS parent, which was consistent with GC‐MS analysis of headspace and solvent‐extracted terpenes showing that cold‐treated GS apples produced higher levels of (E,E)‐α‐farnesene than RG. Transgenic RG fruit downregulated for MdAFS1 expression produced significantly lower levels of (E,E)‐α‐farnesene. To evaluate the role of (E,E)‐α‐farnesene in fungal pathogenesis, MdAFS1 RNA interference transgenic fruit and RG controls were inoculated with three important apple post‐harvest pathogens [Colletotrichum acutatum, Penicillium expansum and Neofabraea alba (synonym Phlyctema vagabunda)]. From results obtained over four seasons, we demonstrate that reduced (E,E)‐α‐farnesene is associated with decreased disease initiation rates of all three pathogens. In each case, the infection rate was significantly reduced 7 days post‐inoculation, although the size of successful lesions was comparable with infections on control fruit. These results indicate that (E,E)‐α‐farnesene production is likely to be an important factor involved in fungal pathogenesis in apple fruit.  相似文献   

17.
The application of helical poly[(S)‐3‐vinyl‐2,2'‐dihydroxy‐1, 1'‐binaphthyl] ( L* ) in the asymmetric borane reduction of prochiral ketones was studied. The results showed that L* had excellent catalytic activity as well as enantioselectivity, giving up to 96% yield and up to 99% enantiomeric excess (ee) of the corresponding secondary alcohol at 25 °C. Moreover, L* can be easily recovered and reused without loss of catalytic activity. Chirality 27:422–424, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

18.
N‐[1‐(4‐(4‐fluorophenyl)‐2,6‐dioxocyclohexylidene)ethyl] (Fde) protected amino acids have been prepared and applied in solid‐phase peptide synthesis monitored by gel‐phase 19F NMR spectroscopy. The Fde protective group could be cleaved with 2% hydrazine or 5% hydroxylamine solution in DMF as determined with gel‐phase 19F NMR spectroscopy. The dipeptide Ac‐L ‐Val‐L ‐Val‐NH2 12 was constructed using Fde‐L ‐Val‐OH and no noticeable racemization took place during the amino acid coupling with N,N′‐diisopropylcarbodiimide and 1‐hydroxy‐7‐azabenzotriazole or Fde deblocking. To extend the scope of Fde protection, the hydrophobic nonapeptide LLLLTVLTV from the signal sequence of mucin MUC1 was successfully prepared using Fde‐L ‐Leu‐OH at diagnostic positions. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
α‐melanocyte stimulating hormone (α‐MSH) is a tridecapeptide fragment of pro‐opiomelanocortin (POMC) with broad effects on appetite, skin pigmentation, hormonal regulation, and potential roles in both inflammation and autoimmunity. The use of this peptide as an anti‐inflammatory agent is limited by its low selectivity between the melanocortin receptors, susceptibility to proteolytic degradation, and rapid clearance from circulation. A retro‐inverso (RI) sequence of α‐MSH was characterized for receptor activity and resistance to protease. This peptide demonstrated surprisingly high selectivity for binding the melanocortin receptor 1 (MC1R). However, RI‐α‐MSH exhibited a diminished binding affinity for MC1R compared to α‐MSH. Mapping of the residues critical for agonist activity, receptor binding, and selectivity by alanine scanning, identified the same critical core tetrapeptide required for the native peptide. Modest improvements in affinity were obtained by conservative changes employing non‐natural amino acids and substitution of the C‐terminal sequence with a portion of a MC1R ligand peptide previously identified by phage display. Recombination of these elements yielded a peptide with an identical Ki as α‐MSH at MC1R and a lower EC50 in Mel‐624 melanoma cells. A number of other structural modifications of the RI peptide were found to differ in effect from those reported for the L ‐form α‐MSH, suggesting a significantly altered interaction with the MC1R. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
David E. Kenny 《Zoo biology》2001,20(4):245-250
After the loss of an African elephant (Loxodonta africana) in February 1989 at the Denver Zoological Gardens (DZG) with very low circulating serum α‐tocopherol, a long‐term study was initiated with three Asian elephants (Elephas maximus) to evaluate the effect of an oral micellized, water‐soluble, natural source d‐α‐tocopherol supplement. Baseline α‐tocopherol levels were evaluated and found to be approximately 3.75‐fold less than those reported for semi‐free‐ranging Asian Nepalese work camp and free‐ranging African elephants. The DZG elephants were then administered a liquid d‐α‐tocopherol (Emcelle®) at 2.2 IU/kg body weight orally once daily. Serum samples were obtained and analyzed at 1, 2, 8, and 12 months and then annually for 96 months. The oral vitamin E supplement significantly elevated serum levels above baseline and were found to be comparable with levels reported for semi–free‐ranging and free‐ranging elephants. Zoo Biol 20:245–250, 2001. © 2001 Wiley‐Liss, Inc.  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