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 31 毫秒
1.
The lipophilicity of a number of N‐acyl derivatives of trans‐ or cis‐: racemic, (1R,2R)‐ or (1S,2S)‐aminocyclohexanol (1–13) exhibiting anticonvulsant activity was investigated. Their lipophilicity (Rm 0) was determined using reversed‐phase thin‐layer chromatography (RP‐TLC) with mixtures of methanol and water as mobile phases. The partition coefficients of compounds 1–13 (log P) were also calculated using two computer programs (Pallas and Chem DU) and compared with Rm 0. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

2.
The lipophilicity of a library of 30 derivatives of dihydrofuran‐2(3H)‐one (γ‐butyrolactone) was determined by MEKC. Calibration curve prepared for ten reference drugs enabled to calculate partition coefficient (log P) for novel compounds. The results of MEKC analysis were compared with lipophilicity coefficients determined by RP‐TLC (RM0) and computational (Mlog P, Clog P) methods. Good correlation was observed between the results obtained by both experimental methods: the MEKC parameters log k and relative lipophilicity RMO. The relationship between determined log P values and results of the computational prediction was weaker. Analysis of the relationship between lipophilicity and anticonvulsant activity showed statistically significant differences between mean values of log P coefficients for group of active (2.18) and inactive (1.51) compounds in the maximal electroshock test.  相似文献   

3.
A number of cinnamamide derivatives possess anticonvulsant activity due to the presence of a number of important pharmacophore elements in their structures. In order to study the correlations between anticonvulsant activity and molecular structure, the crystal structures of three new cinnamamide derivatives with proven anticonvulsant activity were determined by X‐ray diffraction, namely (R,S)‐(2E)‐N‐(2‐hydroxybutyl)‐3‐phenylprop‐2‐enamide–water (3/1), C13H17NO2·0.33H2O, ( 1 ), (2E)‐N‐(1‐hydroxy‐2‐methylpropan‐2‐yl)‐3‐phenylprop‐2‐enamide, C13H17NO2, ( 2 ), and (R,S)‐(2E)‐N‐(1‐hydroxy‐3‐methyl‐butan‐2‐yl)‐3‐phenylprop‐2‐enamide, C14H19NO2, ( 3 ). Compound ( 1 ) crystallizes in the space group P with three molecules in the asymmetric unit, whereas compounds ( 2 ) and ( 3 ) crystallize in the space group P21/c with one and two molecules, respectively, in their asymmetric units. The carbonyl group of ( 2 ) is engaged in an intramolecular hydrogen bond with the hydroxy group. This type of interaction is observed for the first time in these kinds of derivatives. A disorder of the substituent at the N atom occurs in the crystal structures of ( 2 ) and ( 3 ). The crystal packing of all three structures is dominated by a network of O—H…O and N—H…O hydrogen bonds, and leads to the formation of chains and/or rings. Furthermore, the crystal structures are stabilized by numerous C—H…O contacts. We analyzed the molecular structures and intermolecular interactions in order to propose a pharmacophore model for cinnamamide derivatives.  相似文献   

4.
A series of superlipophilic or highly lipophilic semisynthetic betulin derivatives was prepared and their relative lipophilicity was measured by reversed‐phase thin‐layer chromatography (RP‐TLC) at different pH values using 1,4‐dioxane–acetate buffer mixtures as mobile phases. Cholesterol, 17β‐estradiol and pure betulin were used as the reference compounds. Linear relationships were found between RM values and 1,4‐dioxane concentrations in the mobile phases. LogP values were also calculated with computer programs ACD/LogP (ChemSketch 11.0, Advanced Chemistry Development Inc.) and ClogP (Daylight Chemical Information Systems Inc.). The empirical and theoretical data were compared, and the RM0 values correlated well with logP. Two of the synthesized betulin derivatives are reported for the first time. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

5.
A cinnamamide (3‐phenylprop‐2‐enamide) core is present in many pharmacologically active compounds. We report three new crystal structures of N‐substituted cinnamamide derivatives which were screened for anticonvulsant activity, namely (R ,S )‐(2E )‐N‐(2‐hydroxypropyl)‐3‐phenylprop‐2‐enamide, C12H15NO2, ( 1 ), (R ,S )‐(2E )‐N‐(1‐hydroxybutan‐2‐yl)‐3‐phenylprop‐2‐enamide, C13H17NO2, ( 2 ), and (2E )‐1‐(4‐hydroxypiperidin‐1‐yl)‐3‐phenylprop‐2‐en‐1‐one, C14H17NO2, ( 3 ). Compounds ( 1 ) and ( 2 ) crystallize in the Pbca space group with one molecule in the asymmetric unit, whereas compound ( 3 ) crystallizes in the P 21/c space group with two molecules in the asymmetric unit. All the crystal structures are stabilized by intermolecular O—H…O hydrogen bonds and additionally by N—H…O hydrogen bonds in the structures of ( 1 ) and ( 2 ). The investigated compounds possess fragments that are considered as beneficial for anticonvulsant activity. The conformations of these compounds were analyzed in comparison with the characteristic features of the proposed pharmacophore model of anticonvulsants active in the maximal electroshock test, i.e. a phenyl ring or other hydrophobic unit, an electron‐donor atom and a hydrogen‐bond acceptor/donor domain. In the reported series, two calculated distances fitted the reference model, while the third did not. Structure–activity analysis suggests that anticonvulsant properties may be related to the N‐atom substituent. It is beneficial to combine an electron‐donor atom (e.g. an O atom) with an H atom in the substituent to ensure appropriate interactions with the molecular target. We analyzed the intermolecular interactions in order to find an appropriate spatial arrangement of the important features responsible for anticonvulsant activity.  相似文献   

6.
7.
The synthesis of 46 derivatives of (2R,3R,4S)‐2‐(aminomethyl)pyrrolidine‐3,4‐diol is reported (Scheme 1 and Fig. 3), and their inhibitory activities toward α‐mannosidases from jack bean (B) and almonds (A) are evaluated (Table). The most‐potent inhibitors are (2R,3R,4S)‐2‐{[([1,1′‐biphenyl]‐4‐ylmethyl)amino]methyl}pyrrolidine‐3,4‐diol ( 3fs ; IC50(B)=5 μM , Ki=2.5 μM ) and (2R,3R,4S)‐2‐{[(1R)‐2,3‐dihydro‐1H‐inden‐1‐ylamino]methyl}pyrrolidine‐3,4‐diol ( 3fu ; IC50(B)=17 μM , Ki=2.3 μM ). (2S,3R,4S)‐2‐(Aminomethyl)pyrrolidine‐3,4‐diol ( 6 , R?H) and the three 2‐(N‐alkylamino)methyl derivatives 6fh, 6fs , and 6f are prepared (Scheme 2) and found to inhibit also α‐mannosidases from jack bean and almonds (Table). The best inhibitor of these series is (2S,3R,4S)‐2‐{[(2‐thienylmethyl)amino]methyl}pyrrolidine‐3,4‐diol ( 6o ; IC50(B)=105 μM , Ki=40 μM ). As expected (see Fig. 4), diamines 3 with the configuration of α‐D ‐mannosides are better inhibitors of α‐mannosidases than their stereoisomers 6 with the configuration of β‐D ‐mannosides. The results show that an aromatic ring (benzyl, [1,1′‐biphenyl]‐4‐yl, 2‐thienyl) is essential for good inhibitory activity. If the C‐chain that separates the aromatic system from the 2‐(aminomethyl) substituent is longer than a methano group, the inhibitory activity decreases significantly (see Fig. 7). This study shows also that α‐mannosidases from jack bean and from almonds do not recognize substrate mimics that are bulky around the O‐glycosidic bond of the corresponding α‐D ‐mannopyranosides. These observations should be very useful in the design of better α‐mannosidase inhibitors.  相似文献   

8.
Two novel and three new sterol glycosides were isolated from the MeOH extract of the aerial parts of Ajuga salicifolia (L.) Schreber . The structures of the compounds were elucidated as (3R,16S,17S,20R,22S,23S, 24S,25S)‐16,23 : 16,27 : 22,25‐triepoxy‐3‐(β‐D ‐glucopyranosyloxy)coprostigmast‐7‐en‐17‐ol ( 1 ), (3R,16S,17S, 20R,22S,23S,24S,25S)‐16,23 : 16,27 : 22,25‐triepoxy‐3‐{[β‐D ‐glucopyranosyl‐(1→2)‐β‐D ‐glucopyranosyl]oxy}coprostigmast‐7‐en‐17‐ol ( 2 ), (3R,16S,17R,20S,22R,24S,25S)‐22,25‐epoxy‐3,27‐bis(β‐D ‐glucopyranosyloxy)coprostigmast‐7‐en‐16‐ol ( 3 ), (3R,16S,17R,20S,22R,24S,25S)‐22,25‐epoxy‐3‐{[β‐D ‐glucopyranosyl‐(1→2)‐β‐D ‐glucopyranosyl]oxy}‐27‐(β‐D ‐glucopyranosyloxy)coprostigmast‐7‐en‐16‐ol ( 4 ), and (3R,16R,17S,20R,22S,23S, 24S,25S)‐22,25‐epoxy‐3‐(β‐D ‐glucopyranosyloxy)coprostigmast‐7‐ene‐16,17,23,27‐tetrol 27‐acetate ( 5 ) by means of 1D and 2D NMR spectroscopy and HR‐MALDI mass spectrometry. The novel compounds, which consist of three additional ring systems at the coprostigmastane skeleton, were named ajugasalicioside A ( 1 ) and B ( 2 ), and the new compounds C ( 3 ), D ( 4 ) and E ( 5 ). In our cytotoxicity assays (HeLa cells, Jurkat T cells, and peripheral mononuclear blood cells), ajugasaliciosides A–D specifically inhibited the viability and growth of Jurkat T‐leukemia cells at concentrations below 10 μM . Ajugasalicioside A ( 1 ; (IC50=6 μM ) and C ( 3 ; IC50=3 μM ) were the most active compounds. Ajugasalicioside A ( 1 ) induced cell‐cell contact, inhibited Jurkat T cell proliferation, and up‐regulated mRNA levels of the cell‐cycle regulator cyclin D1, which might be an indication for cell differentiation. Furthermore, 1 down‐regulated the mRNA levels of the NF‐κB subunit p65 in a concentration‐dependent manner. These effects were not found for ajugasalicioside B ( 2 ), which has an additional glucose unit, and the onset of cytotoxicity of 2 (IC50=10 μM ) was delayed by 24 h.  相似文献   

9.
The chemoenzymatic synthesis of a collection of pyrrolidine‐type iminosugars generated by the aldol addition of dihydroxyacetone phosphate (DHAP) to C‐α‐substituted N‐Cbz‐2‐aminoaldehydes derivatives, catalyzed by DHAP aldolases is reported. L ‐Fuculose‐1‐phosphate aldolase (FucA) and L ‐rhamnulose‐1‐phosphate aldolase (RhuA) from E. coli were used as biocatalysts to generate configurational diversity on the iminosugars. Alkyl linear substitutions at C‐α were well tolerated by FucA catalyst (i.e., 40–70 % conversions to aldol adduct), whereas no product was observed with C‐α‐alkyl branched substitutions, except for dimethyl and benzyl substitutions (20 %). RhuA was the most versatile biocatalyst: C‐α‐alkyl linear groups gave the highest conversions to aldol adducts (60–99 %), while the C‐α‐alkyl branched ones gave moderate to good conversions (50–80 %), with the exception of dimethyl and benzyl substituents (20 %). FucA was the most stereoselective biocatalyst (90–100 % anti (3R,4R) adduct). RhuA was highly stereoselective with (S)‐N‐Cbz‐2‐aminoaldehydes (90–100 % syn (i.e., 3R,4S) adduct), whereas those with R configuration gave mixtures of anti/syn adducts. For iPr and iBu substituents, RhuA furnished the anti adduct (i.e., FucA stereochemistry) with high stereoselectivity. Molecular models of aldol products with iPr and iBu substituents and as complexes with the RhuA active site suggest that the anti adducts could be kinetically preferred, while the syn adducts would be the equilibrium products. The polyhydroxylated pyrrolidines generated were tested as inhibitors against seven glycosidases. Among them, good inhibitors of α‐L ‐fucosidase (IC50=1–20 μM ), moderate of α‐L ‐rhamnosidase (IC50=7–150 μM ), and weak of α‐D ‐mannosidase (IC50=80–400 μM ) were identified. The apparent inhibition constant values (Ki) were calculated for the most relevant inhibitors and computational docking studies were performed to understand both their binding capacity and the mode of interaction with the glycosidases.  相似文献   

10.
The chromatographic behaviour of the parabens has been investigated on RP‐18F254S, RP‐18WF254S, CNF254S, Diol F254s and silica gel 60F254 plates impregnated with different oils (paraffin, olive, sunflower and corn) using methanol–water mixtures in different volume proportions as mobile phases, the regression determination coefficients being excellent (higher than 0.98 for the majority of compounds). Moreover, highly significant correlations were obtained between different experimental indices of lipophilicity (RM0, b and scores corresponding to the first principal component (PC1)) and computed log P values. All types of stationary phases investigated appear to be highly suited for estimating the lipophilicity of the parabens.  相似文献   

11.
The crystal structures of salt 8 , which was prepared from (R)‐2‐methoxy‐2‐(2‐naphthyl)propanoic acid ((R)‐MβNP acid, (R)‐ 2 ) and (R)‐1‐phenylethylamine ((R)‐PEA, (R)‐ 6 ), and salt 9 , which was prepared from (R)‐2‐methoxy‐2‐(1‐naphthyl)propanoic acid ((R)‐MαNP acid, (R)‐ 1 ) and (R)‐1‐(p‐tolyl)ethylamine ((R)‐TEA, (R)‐ 7 ), were determined by X‐ray crystallography. The MβNP and MαNP anions formed ion‐pairs with the PEA and TEA cations, respectively, through a methoxy‐group‐assisted salt bridge and aromatic CH???π interactions. The networks of salt bridges formed 21 columns in both salts. Finally, (S)‐(2E,6E)‐(1‐2H1)farnesol ((S)‐ 13 ) was prepared from the reaction of (2E,6E)‐farnesal ( 11 ) with deuterated (R)‐BINAL‐H (i.e., (R)‐BINAL‐D). The enantiomeric excess of compound (S)‐ 13 was determined by NMR analysis of (S)‐MαNP ester 14 . The solution‐state structures of MαNP esters that were prepared from primary alcohols were also elucidated.  相似文献   

12.
The lipophilicity (RM0) and specific hydrophobic surface area for the representatives of four generation cephalosporins have been determined by reversed‐phase thin‐layer chromatography, and the effect of different mobile‐phase modifiers (such as methanol, acetonitrile, acetone, 1,4‐dioxane and 2‐propanol) on the retention has been studied. The compounds studied showed typical retention behavior; their RM values decreased linearly with increasing concentration of the organic modifier in the eluent. The linear correlations between the volume fraction of the organic solvent and the RM values over a limited range were established for each solute, resulting in high values of correlation coefficients (>0.95 in most cases). RM values were determined by various concentrations of organic modifier, and the correlation obtained was extrapolated to 0% of organic modifier. Chromatographically established logP (RM0) parameters were compared with computationally calculated partition coefficients values (AClogP, ALOGP, KOWWIN, ALOGPs, XLOGP2, MLOGP and XLOGP3) and experimental octanol–water logP values (measured by the shake flask method). The received results demonstrate that RP‐TLC may be a good alternative technique for analytics in describing the lipophilic nature of investigated cephalosporins as well as the activity. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

13.
Cytotoxicity against the KB cancer cell line as a lead bioactivity‐guided fractionation of the petroleum ether extract of rhizomes of Amomum aculeatum Roxb. led to the isolation of three novel dioxadispiro[5.1.5.2]pentadeca‐9,12‐dien‐11‐one derivatives. The structures of aculeatin A ( 1 ), aculeatin B ( 2 ), and aculeatin C ( 3 ) were established as rel‐(2R,4R,6S)‐ and rel‐(2R,4R,6R)‐4‐hydroxy‐2‐tridecyl‐1,7‐dioxadispiro[5.1.5.2]pentadeca‐9,12‐dien‐11‐one ( 1 and 2 , resp.) and rel‐(2R,4R,6R)‐2‐[4‐(3‐dodecyl‐2‐heptyl‐3‐hydroxy‐6‐oxocylohexa‐1,4‐dienyl)‐2‐oxobutyl]‐4‐hydroxy‐1,7‐dioxadispiro[5.1.5.2]pentadeca‐9,12‐dien‐11‐one ( 3 ) by extensive spectroscopic analyses, particularly 13C‐NMR, inverse‐gated 13C, HMQC, HMBC, NOESY, and INADEQUATE NMR experiments as well as mass spectrometry. The aculeatins represent a novel type of natural products. All compounds showed high cytotoxicity against the KB cell line: 1 , IC50=1.7 μM ; 2 , IC50=2.0 μM ; 3 , IC50=1.6 μM . Additional testing against two Plasmodium falciparum strains as well as against trypomastigote forms of Trypanosoma brucei rhodesiense and Trypanosoma cruzi showed strong activities, particularly against P. falciparum strain K1 ( 1 , IC50=0.18 μM ; 2 , IC50=0.43 μM ; 3 , IC50=0.37 μM ).  相似文献   

14.
15.
p‐(2,2,3,3‐Tetracyanocyclopropyl)phenoxyethyl acrylate ( 5a ) and p‐(2,2,3‐tetracyanocyclopropyl)phenoxyethyl methacrylate ( 5b ) were prepared by the reactions of bromomalononitrile with p‐(2‐acryloyloxyethoxy)benzylidenemalononitrile and p‐(2‐methacryloyloxyethoxy)benzylidenemalononitrile, respectively. Monomers 5a and 5b were polymerized with free‐radical initiators to obtain polymers with multicyanocyclopropane functionalities in the pendant group. The resulting polymers were soluble in acetone, and the inherent viscosities were 0.25–0.30 dL/g. Solution‐cast films showed thermal stability up to 300 °C with glass‐transition temperatures of 140–156 °C. The dipole moments of 5a and 5b , calculated by the atom superposition and electron delocalization molecular orbital method, were 7.58–7.30 D. Piezoelectric coefficients (d31) of the poled polymer films were 1.8–1.9 pC/N, acceptable values for piezoelectric device applications. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 40: 379–384, 2002  相似文献   

16.
The chromatographic behaviour of salicylic acid derivatives was investigated using reversed‐phase high performance thin‐layer chromatography (RP HPTLC) with methanol–water and dioxane–water binary mixtures as mobile phase in order to establish relationships between chromatographic data and selected physico‐chemical parameters that are related to ADME (absorption, distribution, metabolism and elimination). Some of the investigated compounds were screened for antioxidant activity. Examination of chromatographic behaviour revealed a linear correlation between RM values and the volume fraction of mobile phase modifier. Obtained RM0 values were correlated with lipophilicity, solubility, human intestinal absorption, plasma‐protein binding, and blood–brain barrier data. The comparison among chromatographic data obtained by two mobile phase was performed with a statistical technique, principle component analysis. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
A series of newly synthesized 1,3‐purinodiones with potential anticonvulsant activity, exhibiting affinity to adenosine A1 and/or A2A receptors, were subjected to micellar LC (MLC) with SDS as micelle‐forming agent and n‐propanol as organic modifier. Two C18 silica‐based columns were employed in MLC: a particle one and a monolithic. In parallel, those derivatives were also analyzed in RP‐LC on four silica‐based columns and on an immobilized artificial membrane column. The correlations between the relevant logarithms of the retention factors of analytes obtained in MLC, immobilized artificial membrane and RP‐LC systems on the one hand, and the calculated log P (clog P) and log D values (clog D) on the other, were examined. The level of the correlations of retention data from MLC and RP‐LC systems with clog P and clog D obtained is similar but it could be stressed that MLC allows increasing the speed of analysis and using only one mobile phase. Moreover, there is no need of applying an extrapolation procedure in lipophilicity determination. Therefore, the MLC systems, providing chromatographic data in a fast and efficient manner, were demonstrated as promising alternatives to the classical RP‐LC systems to estimate the lipophilicity of drugs and drug candidates.  相似文献   

18.
Nineteen new complexes of carboxylates with transition and rare elements as central ions and their ligands were characterized by chromatographic analyses. The parameter of relative lipophilicity (R M0) of the tested compounds was determined experimentally by the reversed‐phase high‐performance thin layer chromatography method with mixtures of various organic modifiers (acetonitrile, acetone, dioxane) and water as a mobile phase. The extrapolated R M0 values were compared with the logP values calculated from the molecular structures of tested solutes. Similarities between the lipophilicity indices were analysed by principal component analysis and linear regression. Thin‐layer chromatography combined with a magnetic field has been proposed as a complementary method for determination of lipophilicity of the investigated compounds. The chromatograms in the field and outside it were developed simultaneously in two identical chromatographic chambers. One of them was placed in the external magnetic field of 0.4 T inductivity. We proved that chelation causes a drastic change in compound lipophilicity, but all complexes did not exhibit enhanced activity as compared with the parent ligand. Also in the magnetic field the retention of some complexes changed, which means that the presence of the field influences the physicochemical properties of the compounds and their interactions with the stationary phase.  相似文献   

19.
The two epimers (?)‐ 1a and (?)‐ 1b of the macrocyclic lactam alkaloid 3‐hydroxycelacinnine with the (2R,3R) and (2R,3S) absolute configurations, respectively, were synthesized by an alternative route involving macrocyclization with the regio‐ and stereoselective oxirane‐ring opening by the terminal amino group (Schemes 2 and 6). Properly N‐protected chiral trans‐oxirane precursors provided (2R,3R)‐macrocycles after a one‐pot deprotection‐macrocyclization step under moderate dilution (0.005–0.01M ). The best yields (65–85%) were achieved with trifluoroacetyl protection. Macrocyclization of the corresponding cis‐oxiranes was unsuccessful for steric reasons. Inversion at OH? C(3) via nucleophilic displacement of the cyclic sulfamidate derivative with NaNO2 led to (2R,3S)‐macrocycles. The synthesized (?)‐(2R,3S)‐3‐hydroxycelacinnine ((?)‐ 1b ) was identical to the natural alkaloid.  相似文献   

20.
The measurement of the apparent elongation viscosity (ηe) of several polyolefin melts was conducted in this study by using the isothermal fiber‐spinning method. The White–Metzner (W–M) model was used to analyze the spinning flow of the polymer melts and, thus, the elongation viscosity was predicted at elongation strain rates ranging from 0 to approximately 5 s?1. The values of the model parameters required in the W–M model were obtained by curve fitting the experimental data obtained from the shear measurements. The elongation viscosity predicted using the W–M model was in good agreement with the experimental results of fiber spinning. In addition, ηe could also be estimated directly from the measured shear viscosity (ηS) with a formulation using the W–M model; the subsequently obtained elongation viscosity and Trouton ratio (TR) were reasonable within a wide range of strain rates. Based on the experimental and theoretical results, the polyolefin with a high molecular weight was observed to have high elongation viscosity, and the polymer with a broad molecular weight distribution also possessed high ηe. The TR value of the commercial polypropylene (PP‐1040) began to increase from 3 at a deformation rate of 0.1 s?1 and grew up asymptotically to 10, whereas the TR of high‐density polyethylene (HDPE‐606) remained nearly at 3 within the entire range of strain rates. Copyright © 2014 John Wiley & Sons, Ltd.  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