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 78 毫秒
1.
Chiral resolutions of trifluoroacetyl‐derivatized 1‐phenylalkylamines with different type and position of substituent were investigated by capillary gas chromatography by using heptakis(2,3‐di‐O‐methyl‐6‐Otert‐butyldimethylsilyl)‐β‐cyclodextrin diluted in OV‐1701 as a chiral stationary phase. The influence of column temperature on retention and enantioselectivity was examined. All enantiomers of meta‐substituted analytes as well as fluoro‐substituted analytes could be resolved. Temperature had a favorable influence on enantioselectivity for small amines with substituents at the ortho‐position. The type of substituent at the stereogenic center of amines also had a crucial effect as the ethyl group led to poor enantioseparation. Among all analytes studied, trifluoroacetyl‐derivatized 1‐(2′‐fluorophenyl)ethylamine exhibited baseline resolution with the shortest analysis time.  相似文献   

2.
In the synthesis performed in this study, derivatives of 4‐tert‐butylcyclohexanone 1 were obtained using typical reactions of organic synthesis. The bioactivity of the selected compounds was evaluated. 1‐(Bromomethyl)‐8‐tert‐butyl‐2‐oxaspiro[4.5]decan‐3‐one ( 5 ) was characterized by attractant properties against larvae and a weak feeding deterrent activity against adults of Alphitobius diaperinus Panzer . This bromolactone was a moderate antifeedant towards Myzus persicae Sulzer . In addition, ethyl (4‐tert‐butylcyclohexylidene)acetate ( 2 ) and bromolactone 5 displayed antibacterial activity. The strongest bacteriostatic effect was observed against Gram‐positive strains: Bacillus subtilis and Staphylococcus aureus. The bromolactone 5 also limited the growth of Escherichia coli strain.  相似文献   

3.
Three tricyclic 1,2‐dioxetane derivatives, 1a, 2a and 3a were synthesized from their corresponding 1,4‐dioxin acenaphthylene compounds, 1, 2 and 3, by reaction with singlet‐oxygen (1O2) in dichloromethane. Evidence for the formation of the dioxetanes 1a, 2a and 3a is provided by the chemiluminescence (CL) that corresponds to the emission from the electronically excited diesters 1b*, 2b* and 3b*, which are decomposed thermally from the dioxetanes 1a, 2a and 3a, respectively. The highly strained 1,2‐dioxetane ring decomposes from a twisted geometry by simultaneous cleavages of the O–O and C–C bonds, producing the electronically excited diester that emits CL. It was observed that the CL from compound 2a is red‐shifted relative to that of compounds 1a and 3a suggesting a higher degree of stabilization for the excited state by the electron‐donating methoxy group. Also, a study of the solvent effect on fluorescence shows a significant red‐shift in compound 2b, indicating a more polar excited state. The kinetics of the thermal decomposition of the 1,2‐dioxetanes clearly demonstrate that the CL characteristics of compound 2a are quite different from those of compounds, 1a and 3a. These results are consistent with the proposed intramolecular chemically initiated electron exchange luminescence (CIEEL) mechanism which is triggered by the electron‐donating group of compound 2a. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

4.
Exposure of the skin to certain phenols or catechols such as 4‐tert‐butylphenol (TBP) and 4‐tert‐butylcatechol (TBC) may cause leukoderma. These substances are used in the polymer industry and numerous cases have been reported. Several theories of the mechanism for chemical leukoderma have been suggested. In the present study, TBP and TBC are shown to be oxidised by tyrosinase. The oxidation of TBC yields a quinone that is further investigated on its reactions with cysteine or glutathione (GSH). The products formed are isolated and identified by mass spectrometry and nuclear magnetic resonance as being 4‐tert‐butyl‐6‐S‐cysteinylcatechol (cys‐TBC) and 4‐tert‐butyl‐6‐S‐glutathionylcatechol (GS‐TBC). The reactive quinone is a strongly electrophilic substance that rapidly reacts with GSH. A depletion of the GSH defence system may give conditions where the quinone lives long enough to effect its toxic properties. The influence of the reactive tert‐butylquinone on enzymatic activities is demonstrated by the inhibition of glyceraldehyde‐3‐phosphate dehydrogenase.  相似文献   

5.
N‐alkylated trans‐diaziridines are an intriguing class of compounds with two stereogenic nitrogen atoms which easily interconvert. In the course of our investigations of the nature of the interconversion process via nitrogen inversion or electrocyclic ring opening ring closure, we synthesized and characterized the three constitutionally isomeric diaziridines 1,2‐di‐n‐propyldiaziridine 1 , 1‐isopropyl‐2‐n‐propyldiaziridine 2 , and 1,2‐diisopropyldiaziridine 3 to study the influence of the substituents on the interconversion barriers. Enantiomer separation was achieved by enantioselective gas chromatography on the chiral stationary phase Chirasil‐β‐Dex with high separation factors α (1‐isopropyl‐2‐n‐propyldiaziridine: 1.18; 1, 2‐diisopropyldiaziridine: 1.24; 100°C 50 kPa He) for the isopropyl substituted diaziridines. These compounds showed pronounced plateau formation between 100 and 150°C, and peak coalescence at elevated temperatures. The enantiomerization barriers ΔG? and activation parameters ΔH? and ΔS? were determined by enantioselective dynamic gas chromatography (DGC) and direct evaluation of the elution profiles using the unified equation implemented in the software DCXplorer. Interestingly, 1‐isopropyl‐2‐n‐propyldiaziridine and 1,2‐diisopropyldiaziridine exhibit similar high interconversion barriers ΔG? (100°C) of 128.3 ± 0.4 kJ mol?1 and 129.8 ± 0.4 kJ mol?1, respectively, which indicates that two sterically demanding substituents do not substantially increase the barrier as expected for a distinct nitrogen inversion process. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

6.
A bicyclic dioxetane 1 bearing a 4‐(4‐cyanophenyl)iminomethyl‐3‐hydroxyphenyl group was found to undergo base‐induced decomposition with the accompanying emission of light, the colour of which changed depending on the base used and its concentration. When 1 was triggered with tetrabutylammonium fluoride (TBAF), 1 displayed an emission of glowing orange light. On the other hand, on treatment with a high concentration of potassium t‐butoxide complexed with 18‐crown‐6 ether, 1 afforded a flash of blue light. The mechanistic study of this unprecedented phenomenon revealed that the emission of glowing orange light was due to the normal oxido anion of keto ester 11, whereas the emission of a flash of blue light was attributed to another species that was produced by addition of a nucleophile to an iminomethyl of unstable oxido anion of dioxetane 10. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

7.
We investigated the stereodynamics of 5,5’‐substituted tropos BIPHEP ligands (2,2’‐bis(diphenylphosphino)‐biphenyls) by enantioselective dynamic high‐performance liquid chromatography (DHPLC) to elucidate the influence of the substitution pattern and electronics of the substituents (methyl, methoxy, and hydroxyl groups). By temperature‐dependent dynamic HPLC measurements the activation parameters ΔG, ΔH, and ΔS could be determined with high precision, revealing that the activation barrier of these 5,5’‐substituted BIPHEP ligands ranges in a narrow band between 87.8 and 93.0 kJ mol–1, making them highly attractive as deracemizable dynamic chiral ligands in asymmetric catalysis. Interestingly, the activation parameters are highly influenced by a hydroxyl or methoxy group in the 5,5’‐position of the BIPHEP ligands. Chirality 25:126–132, 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

8.
Bicyclic dioxetanes 2a–c bearing a 2′‐alkoxy‐2‐hydroxy‐1,1′‐binaphthyl‐7‐yl moiety were effectively synthesized and their base‐induced chemiluminescent decomposition was investigated by the use of alkaline metal (Na+ and K+) or Mg2+ alkoxide in MeOH. When 2a–c were treated with tetrabutylammonium fluoride (TBAF) in dimethyl sulfoxide (DMSO) as a reference system, they showed chemiluminescence as a flash of orange light (maximum wavelength λmaxCL = 573–577 nm) with efficiency ΦCL = 6–8 × 10–2. On the other hand, for an alkaline metal (Na+ or K+) alkoxide/MeOH system, 2a–c decomposed slowly to emit a glow of chemiluminescence, the spectra of which were shifted slightly toward red from the TBAF/DMSO system, and ΦCL (= 1.4–2.3 × 10–3) was considerably decreased. In addition, Mg(OMe)2 was found to play a characteristic role as a base for the chemiluminescent decomposition of 2a–c through coordination to the intermediary oxidoaryl‐substituted dioxetanes 13. Thus, Mg2+ increased ΦCL to more than twice those with Na+ or K+, while it shifted λmaxCL considerably toward blue (λmaxCL = 550–566 nm). Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

9.
A colorimetric and turn‐on fluorescent probe for fluoride ions, tert‐butyldimethylsilane 6‐acetyl‐2‐naphtholate, was readily synthesized from 6‐acetyl‐2‐naphthol and tert‐butyldimethylchlorosilane (TBSCl). The probe exhibits high sensitivity and good selectivity for fluoride ions in acetonitrile. The inherent mechanism involves the cleavage of the Si–O bond in the probe, which induced yellow color formation and prominent fluorescence enhancement. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

10.
Three sodium salts of (2E)‐3‐(4'‐halophenyl)prop‐2‐enoyl sulfachloropyrazine (CCSCP) were synthesized and their structures were determined by 1H and 13C NMR, LC‐MS and IR. The binding properties between CCSCPs and bovine serum albumin (BSA) were studied using fluorescence spectroscopy in combination with UV–vis absorbance spectroscopy. The results indicate that the fluorescence quenching mechanisms between BSA and CCSCPs were static quenching at low concentrations of CCSCPs or combined quenching (static and dynamic) at higher CCSCP concentrations of 298, 303 and 308 K. The binding constants, binding sites and corresponding thermodynamic parameters (ΔH, ΔS, ΔG) were calculated at different temperatures. All ΔG values were negative, which revealed that the binding processes were spontaneous. Although all CCSCPs had negative ΔH and positive ΔS, the contributions of ΔH and ΔS to ΔG values were different. When the 4'‐substituent was fluorine or chlorine, van der Waals interactions and hydrogen bonds were the main interaction forces. However, when the halogen was bromine, ionic interaction and proton transfer controlled the overall energetics. The binding distances between CCSCPs and BSA were determined using the Förster non‐radiation energy transfer theory and the effects of CCSCPs on the conformation of BSA were analyzed by synchronous fluorescence spectroscopy. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

11.
The reduction of the axially chiral N‐(o‐aryl)‐5,5‐dimethyl‐2,4‐oxazolidinediones by NaBH4 yielded axially chiral N‐(o‐aryl)‐4‐hydroxy‐5,5‐dimethyl‐2‐oxazolidinone enantiomers having a chiral center at C‐4, with 100% diastereoselectivity as has been shown by their 1H and 13C NMR spectra and by enantioselective HPLC analysis. The resolved enantiomeric isomers were found to interconvert thermally through an aldehyde intermediate formed upon ring cleavage via a latent ring‐chain‐ring tautomerization. It was found that the rate of enantiomerization depended on the size and the electronic effect of the ortho substituent present on the aryl ring bonded to the nitrogen of the heterocycle. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

12.
The influence of side‐chain structure on the mode of reaction of ortho‐quinone amines has been investigated with a view, ultimately, to developing potential methods of therapeutic intervention by manipulating the early stages of melanogenesis. Four N‐substituted dopamine derivatives have been prepared and quinone formation studied using pulse radiolysis and tyrosinase‐oximetry. Ortho‐quinones with an amide or urea side chain were relatively stable, although evidence for slow formation of isomeric para‐quinomethanes was observed. A thiourea derivative cyclized fairly rapidly (k = 1.7/s) to a product containing a seven‐membered ring, whereas a related amidine gave more rapidly (k ~ 2.5 × 102/s) a stable spirocyclic product. The results suggest that cyclization of amides, ureas and carbamates (NHCO‐X; X = R, NHR or OR) does not occur and is not, therefore, a viable approach to the formation of tyrosinase‐activated antimelanoma prodrugs. It is also concluded that for N‐acetyldopamine spontaneous ortho‐quinone to para‐quinomethane isomerization is slow.  相似文献   

13.
In order to assign the absolute configurations of 8‐tert‐butyl‐2‐hydroxy‐7‐methoxy‐8‐methyl‐9‐oxa‐6‐azaspiro[4.5]dec‐6‐en‐10‐one ( 2a , 2b ), their esters ( 5a , 5b , 5c , 5d ) with (R)‐ or (S)‐2‐methoxyphenylacetic acid ( 4a , 4b ) have been synthesized. The absolute configurations of these compounds have been determined on the basis of NOESY correlations between the protons of the tert‐butyl group and the cyclopentane fragment of the molecules. The crucial part of this analysis was assignment of the absolute configuration at C‐5. Additionally, by calculation of the chemical shift anisotropy, δRS, for the relevant protons, it was also possible to confirm the absolute configurations at the C‐2 centres of compounds 2a , 2b and 5a , 5b , 5c , 5d . Chirality, 25:422–426, 2013.© 2013 Wiley Periodicals, Inc.  相似文献   

14.
We have reported that benzene‐1,2‐, 1,3‐, and 1,4‐di‐N‐substituted carbamates ( 1–15 ) are characterized as the conformationally constrained inhibitors of acetylcholinesterase and mimic gauche, eclipsed, and anti‐conformations of acetylcholine, respectively (J Biochem Mol Toxicol 2007;21:348–353). We further report the inhibition of butyrylcholinesterase by these inhibitors. Carbamates 1–15 are also characterized as the pseudosubstrate inhibitors of butyrylcholinesterase as in the acetylcholinesterase catalysis. Benzene‐1,4‐di‐Nn‐hexylcarbamate ( 12 ) and benzene‐1,4‐di‐Nn‐octylcarbamate ( 13 ) are the two most potent inhibitors of butyrylcholinesterase among inhibitors 1–15 . These two para compounds, with the angle of 180° between two C(benzene)? O bonds, mimic the preferable anti C? O/C? N conformers for the choline ethylene backbone of butyrylcholine during the butyrylcholinesterase catalysis. The second n‐hexylcarbamyl or n‐octylcarbamyl moiety of inhibitors 12 and 13 is proposed to bind tightly to the peripheral anionic site of butyrylcholinesterase from molecular modeling. Butyrylcholinesterase prefers para‐carbamates to ortho‐ and meta‐carbamates, whereas acetylcholinesterase prefers para‐ and meta‐carbamates to ortho‐carbamates. This result implies that the anionic site of butyrylcholinesterase is relatively smaller than that of acetylcholinesterase because meta‐carbamates, which may bind to the anionic sites of both enzymes, are not potent inhibitors of butyrylcholinesterase. © 2009 Wiley Periodicals, Inc. J Biochem Mol Toxicol 23:303–308, 2009; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/jbt.20286  相似文献   

15.
16.
Young Kee Kang  In Kee Yoo 《Biopolymers》2014,101(11):1077-1087
Conformational preferences of 9‐ and 14‐helix foldamers have been studied for γ‐dipeptides of 2‐aminocyclohexylacetic acid (γAc6a) residues such as Ac‐(γAc6a)2‐NHMe ( 1 ), Ac‐(Cα‐Et‐γAc6a)2‐NHMe ( 2 ), Ac‐(γAc6a)2‐NHBn ( 3 ), and Ac‐(Cα‐Et‐γAc6a)2‐NHBn ( 4 ) at the M06‐2X/cc‐pVTZ//M06‐2X/6‐31 + G(d) level of theory to explore the influence of substituents on their conformational preferences. In the gas phase, the 9‐helix foldamer H9 and 14‐helix foldamer H14‐z are found to be most preferred for dipeptides 2 and 4 , respectively, as for dipeptides 1 and 3 , which indicates no remarkable influence of the Cα‐ethyl substitution on conformational preferences. The benzyl substitution at the C‐terminal end lead H14‐z to be the most preferred conformer for dipeptides 3 and 4 , whereas it is H9 for dipeptides 1 and 2 , which can be ascribed to the favored C? H···π interactions between the cyclohexyl group of the first residue and the C‐terminal benzyl group. There are only marginal changes in backbone structures and the distances and angles of H‐bonds for all local minima by Cα‐ethyl and/or benzyl substitutions. Although vibrational frequencies and intensities of the dipeptide 4 calculated at both M06‐2X/6‐31 + G(d) and M05‐2X/6‐31 + G(d) levels of theory are consistent with observed results in the gas phase, H14‐z is predicted to be most preferred by ΔG only at the former level of theory. Hydration did not bring the significant changes in backbone structures of helix foldamers for both dipeptide 1 and 4 . It is expected that the different substitutions at the C‐terminal end lead to the different helix foldamers, which may increase the resistance of helical structures to proteolysis and provide the more surface to the helical structures suitable for molecular recognition. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1077–1087, 2014.  相似文献   

17.
The aim of this study was to evaluate the impact that 6‐O‐(3″, 4″‐di‐Otrans‐cinnamoyl)‐α‐ l ‐rhamnopyranosylcatalpol (Dicinn) and verbascoside (Verb), two compounds simultaneously reported in Verbascum ovalifolium, have on tumor cell viability, apoptosis, cell cycle kinetics, and intracellular reactive oxygen species (ROS) level. At 100 µg/mL and 48 hours incubation time, Dicinn and Verb produced good cytotoxic effects in A549, HT‐29, and MCF‐7 cells. Dicinn induced cell‐cycle arrest at the G0/G1 phase and apoptosis, whereas Verb increased the population of subG1 cells and cell apoptosis rates. Furthermore, the two compounds exhibited time‐dependent ROS generating effects in tumor cells (1‐24 hours). Importantly, no cytotoxic effects were induced in nontumor MCF‐10A cells by the two compounds up to 100 µg/mL. Overall, the effects exhibited by Verb in tumor cells were more potent, which can be correlated with its structural features, such as the presence of phenolic hydroxyl groups.  相似文献   

18.
Dehydroamino acids are non‐coded amino acids that offer unique conformational properties. Dehydrophenylalanine (ΔPhe) is most commonly used to modify bioactive peptides to constrain the topography of the phenyl ring in the side chain, which commonly serves as a pharmacophore. The Ramachandran maps (in the gas phase and in CHCl3 mimicking environments) of ΔPhe analogues with methyl groups at the β position of the side chain as well as at the C‐terminal amide were calculated using the B3LYP/6‐31 + G** method. Unexpectedly, β‐methylation alone results in an increase of conformational freedom of the affected ΔPhe residue. However, further modification by introducing an additional methyl group at C‐terminal methyl amide results in a steric crowding that fixes the torsion angle ψ of all conformers to the value 123°, regardless of the Z or E position of the phenyl ring. The number of conformers is reduced and the accessible conformational space of the residues is very limited. In particular, (Z)‐Δ(βMe)Phe with the tertiary C‐terminal amide can be classified as the amino acid derivative that has a single conformational state as it seems to adopt only the β conformation. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
The interaction between 3‐spiro‐2′‐pyrrolidine‐3′‐spiro‐3″‐piperidine‐2,3″‐dione (PPD) and bovine serum albumin (BSA) in aqueous solution was studied using fluorescence and UV–vis spectroscopy. Fluorescence emission data revealed that BSA (1.00 × 10‐5 mol/L) fluorescence was statically quenched by PPD at various concentrations, which implies that a PPD–BSA complex was formed. The binding constant (KA), the number of binding sites (n) and the specific binding site of the PPD with BSA were determined. Energy‐transfer efficiency parameters were determined and the mechanism of the interaction discussed. The thermodynamic parameters, ΔG, ΔH and ΔS, were obtained according to van't Hoff's equation, showing the involvement of hydrophobic forces in these interactions. The effect of PPD acting on the BSA conformation was detected by synchronous fluorescence. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

20.
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   

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