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.
(−)‐ and (+)‐Conduramine B‐1 ((−)‐ and (+)‐ 5 , resp.) have been derived from (+)‐ and (−)‐7‐oxabicyclo[2.2.1]hept‐5‐en‐2‐one (‘naked sugars’ of the first generation). Although (−)‐ 5 imitates the structure of β‐glucosides, it does not inhibit β‐glucosidases but inhibits α‐mannosidases selectively. N‐Benzylation of (−)‐ 5 improves the potency of conduramine B‐1 as α‐mannosidase inhibitor and also generates compounds inhibiting β‐glucosidases. For instance, (−)‐N‐benzyl‐conduramine B‐1 ((−)‐ 19a ) is a competitive inhibitor of β‐glucosidase from almonds (IC50 = 32 μM , Ki = 10 μM ) and a weak inhibitor of α‐mannosidases from jack bean (IC50 = 171 μM ) and from almonds (IC50 = 225 μM ) whereas (−)‐N‐(4‐phenylbenzyl)conduramine B‐1 ((−)‐ 19g ) is a good inhibitor of α‐mannosidase from jack beans (IC50 = 29 μM , Ki = 4.8 μM ) and a weaker inhibitor of β‐glucosidase from almonds (IC50 = 32 μM , Ki = 7.8 μM ) (Table 1).  相似文献   

2.
The gluco‐ and manno‐tetrahydropyridoimidazole‐2‐acetates and ‐acetic acids 16 and 17 , and 20 and 21 , respectively, were synthesized by condensation, in the presence of HgCl2, of the known thionolactam 26 with the β‐amino ester 25 that was obtained by addition of AcOMe to the imine 22 , followed by debenzylation. The resulting methyl esters 16 and 20 were hydrolyzed to the acetic acids 17 and 21 . The (methoxycarbonyl)‐imidazole 14 and the acid 15 were obtained via the known aldehyde 29 . The imidazoles 14 – 17, 20 , and 21 were tested as inhibitors of the β‐glucosidase from Caldocellum saccharolyticum, the α‐glucosidase from brewer's yeast, the β‐mannosidase from snail, and the α‐mannosidase from Jack beans (Tables 1–3). There is a similar dependence of the Ki values on the nature of the C(2)‐substituent in the gluco‐ and manno‐series. With the exception of 19 , manno‐imidazoles are weaker inhibitors than the gluco‐analogues. The methyl acetates 16 and 20 are 3–4 times weaker than the methyl propionates 5 and 11 , in agreement with the hydrophobic effect. The gluco‐configured (methoxycarbonyl)‐imidazole 14 is 20 times weaker than the methyl acetate 16 , reflecting the reduced basicity of 14 , while the manno‐configured (methoxycarbonyl)‐imidazole 18 is only 1.2 times weaker than the methyl acetate 20 , suggesting a binding interaction of the MeOCO group and the β‐mannosidase. The carboxylic acids 6, 12, 15, 17, 19 , and 21 are weaker inhibitors than the esters, with the propionic acids 6 and 12 being the strongest and the carboxy‐imidazoles 15 and 19 the weakest inhibitors. The manno‐acetate 21 inhibits the β‐mannosidase ca. 8 times less strongly than the propionate 12 , but only 1.5 times more strongly than the carboxylate 19 , suggesting a compensating binding interaction also of the COOH group and the β‐mannosidase. The α/β selectivity for the gluco‐imidazoles ranges between 110 for 15 and 13.4?103 for 6 ; the manno‐imidazoles are less selective. The methyl propionates proved the strongest inhibitors of the α‐glucosidase (IC50 ( 5 )=25 μM ) and the α‐mannosidase (Ki( 11 ) =0.60 μM ).  相似文献   

3.
The racemic gluco‐configured norbornanes 4 and 16 were prepared and tested as inhibitors of β‐glucosidases. The known alcohol 5 was deprotected to provide the triol 6 . Silylation (→ 7 ), monobenzoylation (→ 8 / 9 ), and oxidation provided the regioisomeric ketones 10 and 11 . Reduction of 10 gave the desired endo‐alcohol 13 , albeit in low yield, while reduction of the isomeric ketone 11 provided mostly the altro‐configured endo‐alcohol 12 . The alcohol 13 was desilylated to 14 . Debenzoylation to 15 followed by hydrogenolytic deprotection gave the amino triol 4 that was reductively aminated to the benzylamine 16 . The amino triols 4 and 16 proved weak inhibitors of the β‐glucosidase from Caldocellum saccharolyticum ( 4 : IC50 = 5.6 mm; 16 : IC50 = 3.3 mm) and from sweet almonds ( 16 : IC50 = 5.5 mm) . A comparison of 4 with the manno‐configured norbornane 3 shows that 3 is a better inhibitor of snail β‐mannosidase than 4 is of β‐glucosidases, in keeping with earlier results suggesting that these β‐glycosidases enforce a different conformational itinerary.  相似文献   

4.
The N‐unsubstituted D ‐arabino‐tetrahydropyridazinone 7 is a micromolar inhibitor of β‐glucosidases from sweet almonds (competitive), Caldocellum saccharolyticum (mixed), yeast α‐glucosidase (competitive), jack bean α‐mannosidase (competitive), and snail β‐mannosidase (competitive). The N‐substituted tetrahydropyridazinones 22 , 24 , and 26 are weak inhibitors of these glycosidases, and so are the dihydropyridazinones 8 and 17 – 19 , where the best inhibition was observed for 8 (Ki=56 μM for jack bean α‐mannosidase). The tetrahydropyridazinones were obtained by reduction of the corresponding dihydropyridazinones with NaCNBH3, and the dihydropyridazinones were prepared by treatment with hydrazine or substituted hydrazines of the known and readily available D ‐threo‐pent‐2‐uluronate 11 .  相似文献   

5.
Three new xanthones, namely huperxanthones A–C ( 1 – 3 , resp.), were obtained from the cultures of Aspergillus versicolor, a fungal endophyte of Huperzia serrata, together with 1,7‐dihydroxy‐8‐(methoxycarbonyl)xanthone‐3‐carboxylic acid ( 4 ), β‐diversonolic acid methyl ester ( 5 ), 4‐hydroxyvertixanthone ( 6 ), and sydowinin B ( 7 ). The structures of the new compounds were established by detailed NMR and MS analysis, especially by 2D‐NMR experiments. All xanthones were evaluated for their effects on α‐glucosidase. Compound 4 exhibited a potent inhibitory activity against α‐glucosidase with an IC50 value of 0.24 mM (vs. 0.38 mM for acarbose). The rest of the compounds showed weak or no activity against α‐glucosidase.  相似文献   

6.
The known glucaro‐1,5‐lactam 8 , its diastereoisomers 9 – 11 , and the tetrahydrotetrazolopyridine‐5‐carboxylates 12 – 14 were synthesised as potential inhibitors of β‐D ‐glucuronidases and α‐L ‐iduronidases. The known 2,3‐di‐O‐benzyl‐4,6‐O‐benzylidene‐D ‐galactose ( 16 ) was transformed into the D ‐galactaro‐ and L ‐altraro‐1,5‐lactams 9 and 11 via the galactono‐1,5‐lactam 21 in twelve steps and in an overall yield of 13 and 2%, respectively. A divergent strategy, starting from the known tartaric anhydride 41 , led to the D ‐glucaro‐1,5‐lactam 8 , D ‐galactaro‐1,5‐lactam 9 , L ‐idaro‐1,5‐lactam 10 , and L ‐altraro‐1,5‐lactam 11 in ten steps and in an overall yield of 4–20%. The anhydride 41 was transformed into the L ‐threuronate 46 . Olefination of 46 to the (E)‐ or (Z)‐alkene 47 or 48 followed by reagent‐ or substrate‐controlled dihydroxylation, lactonisation, azidation, reduction, and deprotection led to the lactams 8 – 11 . The tetrazoles 12 – 14 were prepared in an overall yield of 61–81% from the lactams 54, 28 , and 67 , respectively, by treatment with Tf2O and NaN3, followed by saponification, esterification, and hydrogenolysis. The lactams 8 – 11 and 40 and the tetrazoles 12 – 14 are medium‐to‐strong inhibitors of β‐D ‐glucuronidase from bovine liver. Only the L ‐ido‐configured lactam 10 (Ki = 94 μM ) and the tetrazole 14 (Ki = 1.3 mM ) inhibit human α‐L ‐iduronidase.  相似文献   

7.
The influence of a substituent at the hydroximo function of the lactam analogue 1 on the inhibition of β- and α-glucosidases is evaluated. In contrast to 1 , the O-alkyl oximes 5 , 6 , 9 , and 10 are selective inhibitors of β-glucosidases. Alkylation of the D -gluconohydroximo-1,5-lactam 19 with the triflate 12 , or condensation of the thiogluconolactam 20 with the hydroxylamines 14 or 18 afforded the benzylated cellobioside analogues 21 and 23 , respectively. The O-alkyl oximes 33 and 39 were prepared similarly (Scheme 3). Deprotection afforded the cellobioside analogues 5 and 6 , and the O-alkyl oximes 9 and 10 . The lactam O-alkyl oximes 5 , 6 , 9 , and 10 are strong inhibitors of the β-glucosidase from C. saccharolyticum (IC50=0.3 – 8 μM ) and, with exception of the dodecyl analogue 9 (IC50=2 μM ), moderate-to-weak inhibitors of β-glucosidases from sweet almond (IC50=60 – 1000 μM ; see Table). In contrast to the strong inhibition of α-glucosidase from brewer's yeast by 1 (Ki=2.9 μM ), the ethers 5 , 6 , and 10 are weak inhibitors of this enzyme (IC50 between 2500 and >5000 μM ). Similarly, the D -galactohydroximo-1,5-lactam 7 is a potent inhibitor of the α-galactosidase from coffee beans and of the β-galactosidases from bovine liver and E. coli (Ki=5, 10, and 0.1 μM , resp.), while the lactoside analogue 8 is a strong inhibitor of the E. coli β-galactosidase (Ki=0.1 μM ), but a moderate-to-weak inhibitor of coffee-bean α-galactosidase and bovine-liver β-galactosidase (Ki=250 μM and IC50=2500 μM , resp.). The galacto-configured lactam oximes 7 and 8 are good inhibitors of the β-glucosidase isolated from C. saccharolyticum (Ki=2.5 and 3.3 μM , resp.).  相似文献   

8.
It was shown that retaining β‐glucosidases and galactosidases of families 1–3 feature a strong interaction between C(2)OH of the substrate and the catalytic nucleophile. An analogous interaction can hardly take place for retaining β‐mannosidases. A structure? activity comparison between the inhibition of the β‐glucosidase from Caldocellum saccharolyticum (family 1) and β‐glucosidase from sweet almonds by the gluco‐imidazoles 1 – 6 , and the inhibition of snail β‐mannosidase by the corresponding manno‐imidazoles 8 – 13 does not show any significant difference, suggesting that also the mechanisms of action of these glycosidases do not differ significantly. For this comparison, we synthesized and tested the manno‐imidazoles 9 – 13, 28, 29, 32, 35, 40, 41, 43, 46, 47 , and 50 . Among these, the alkene 29 is the strongest known inhibitor of snail β‐mannosidase (Ki=6 nM , non‐competitive); the aniline 35 is the strongest competitive inhibitor (Ki=8 nM ).  相似文献   

9.
Seventeen flavonoids, five of which are flavone C‐diosides, 1 – 5 , were isolated from the BuOH‐ and AcOEt‐soluble fractions of the leaf extract of Machilus konishii. Among 1 – 5 , apigenin 6‐Cβ‐D ‐xylopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 2 ), apigenin 8‐Cα‐L ‐arabinopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 4 ), and apigenin 8‐Cβ‐D ‐xylopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 5 ) are new. Both 4 and 5 are present as rotamer pairs. The structures of the new compounds were elucidated on the basis of NMR‐spectroscopic analyses and MS data. In addition, the 1H‐ and 13C‐NMR data of apigenin 6‐Cα‐L ‐arabinopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 3 ) were assigned for the first time. The isolated compounds were assayed against α‐glucosidase (type IV from Bacillus stearothermophilus). Kaempferol 3‐O‐(2‐β‐D ‐apiofuranosyl)‐α‐L ‐rhamnopyranoside ( 12 ) was found to possess the best inhibitory activity with an IC50 value of 29.3 μM .  相似文献   

10.
The gluco‐configured C(2)‐substituted tetrahydroimidazopyridines 8 – 14 were prepared and tested as inhibitors of the β‐glucosidases from Caldocellum saccharolyticum and from sweet almonds, and of the α‐glucosidase from brewer's yeast. All new imidazopyridines are nanomolar inhibitors of the β‐glucosidases and micromolar inhibitors of the α‐glucosidase. The 3‐phenylpropyl derivative 14 proved the strongest inhibitor of the Caldocellum β‐glucosidase (Ki = 0.9 nM ), only slightly weaker than the known 2‐phenylethyl analogue 7 , and the propyl derivative 13 is the strongest inhibitor of the sweet almond β‐glucosidases (Ki = 3.2 nM ), again slightly weaker than 7 . There is no strong dependence of the inhibition on the nature of the C(2)‐substituent and no clear correlation between the inhibitory strength of the known manno‐configured imidazopyridines 2 – 6 and the gluco‐analogues 8 – 12 . While most manno‐imidazopyridines are competitive inhibitors, the gluco‐analogues proved non‐competitive inhibitors of the Caldocellum β‐glucosidase and mixed‐type or partial mixed‐type inhibitors of the sweet almond β‐glucosidases.  相似文献   

11.
The gluco‐configured analogue 15 of nagstatin ( 1 ) and the methyl ester 14 were synthesized via condensation of the thionolactams 17 or 18 with the β‐amino ester 19 . The silyl ethers 20 and 21 resulting from 17 were desilylated to 22 and 23 ; these alcohols were directly obtained by condensing 18 and 19 . The attempted substitution of the C(8)? OH group of 22 by azide under Mitsunobu conditions led unexpectedly to the deoxygenated α‐azido esters 24 . The desired azide 25 was obtained by treating the manno‐configured alcohol 23 with diphenyl phosphorazidate. The azide was transformed to the debenzylated acetamido ester 14 that was hydrolyzed to the nagstatin analogue 15 . The imidazole‐2‐acetates 14 and 15 are nanomolar inhibitors of the N‐acetyl‐β‐glucosaminidases from Jack beans and from bovine kidney, submicromolar to micromolar inhibitors of the β‐glucosidase from Caldocellum saccharolyticum, and rather weak inhibitors of the snail β‐mannosidase. In all cases, the ester was a stronger inhibitor than the corresponding acid. As expected from their gluco‐configuration, both imidazopyridines 14 and 15 are stronger inhibitors of the βN‐acetylglucosaminidase from bovine kidney than nagstatin.  相似文献   

12.
Synthesis of indolyl linked benzylidene based meta‐substituted phenyl containing thiazolidinediones ( 4a – b ), rhodanine ( 5a – b ), and 1,3‐dicarbonyl based acyclic analogs of isoxazolidinediones ( 6a – 7b ) in an effort to develop novel α‐glucosidase inhibitors in the management of hyperglycemia for the treatment of type 2 diabetes is reported. The structure of all the novel synthesized compounds was confirmed through the spectral studies (LC–MS, 1H‐NMR, 13C‐NMR, and FTIR). Comparative evaluation of these compounds revealed that the compound 5b showed maximum inhibitory potential against α‐amylase and α‐glucosidase giving an IC50 value of 0.28 ± 0.01 μM. Furthermore, binding affinities in terms of G score values and hydrogen bond interactions between all the synthesized compounds and the AA residues in the active site of the protein (PDB code: 3TOP) to that of Acarbose (standard drug) were explored with the help of molecular docking studies. Compound 5b was considered as promising candidate of this series.  相似文献   

13.
Three new alkaloids, 2′‐Oβ‐D ‐glucopyranosyl‐11‐hydroxyvincoside lactam ( 1 ), 22‐O‐demethyl‐22‐Oβ‐D ‐glucopyranosylisocorynoxeine ( 2 ), and (4S)‐corynoxeine N‐oxide ( 3 ) were isolated from the leaves of Uncaria rhynchophylla, together with four known tetracyclic oxindole or indole alkaloids, isocorynoxeine N‐oxide ( 4 ), rhynchophylline N‐oxide ( 5 ), isorhynchophylline N‐oxide ( 6 ), and dihydrocorynantheine ( 7 ), and an indole alkaloid glycoside, strictosidine ( 8 ). The structures of 1 – 3 were elucidated by spectroscopic methods including UV, IR, ESI‐TOF‐MS, 1D‐ and 2D‐NMR, as well as CD experiments. The activity assay showed that 8 (IC50=8.3 μM ) exhibited potent inhibitory activity on lipopolysaccharide(LPS)‐induced nitrogen monoxide (NO) release in N9 microglia cells. However, only weak inhibitory activities were observed for 1 – 7 (IC50>100 μM for 1 – 6 or >30 μM for 7 ).  相似文献   

14.
The spirodiaziridines 6 and 9 , potential inhibitors of α‐ and β‐glucosidases, were prepared from the validoxylamine A‐derived cyclohexanone 5 . The trimethylsilyl protecting groups of 5 are crucial for the formation of 6 in good yields. Oxidation of 6 gave 7 . The diaziridine 6 (pKHA=2.6) and the diazirine 7 did not inhibit the β‐glucosidases from almonds, the β‐glucosidase from Caldocellum saccharolyticum, and the α‐glucosidase from yeast. The N‐benzyl diaziridine 9 is a very weak inhibitor of the α‐glucosidase, but did not inhibit the β‐glucosidases. To see whether the weak inhibition is due to the low basicity of the diaziridines or to geometric factors, we prepared the spiro‐aziridines 21 and 25 and 1‐epivalidamine ( 32 ). The known cyclohexanone 10 was methylenated and epoxidised to 16 and 17 . Azide opening of 16 and 17 , mesylation, LiAlH4 reduction, and deprotection gave the aziridines 21 and 25 respectively. 1‐Epivalidamine ( 32 ) was prepared from the known carba‐glucose 29 . The aziridine 25 (pKHA=6.8) is a weak irreversible inhibitor of the β‐glucosidase from Caldocellum saccharolyticum and a weak reversible inhibitor of the α‐glucosidase from yeast, but did not inhibit the β‐glucosidases from almonds. The poorly stable aziridine 21 weakly inhibited the three enzymes. Similarly, 1‐epivalidamine (pKHA=8.4) proved only a weak inhibitor. The known cyclopentylamine 34 (pKHA=7.9), however, is a micromolar inhibitor of these enzymes. The much stronger inhibition by 34 is related to the pseudoaxial orientation of its amino group.  相似文献   

15.
Addition reactions of acid chlorides with various 2‐substituted 4,5‐dihydro‐4,4‐dimethyl‐5‐(methylsulfanyl)‐1,3‐thiazoles under basic conditions were studied. Two kinds of products were obtained from these additions, β‐lactams and non‐β‐lactam adducts. When the reaction was carried out with 4,5‐dihydro‐1,3‐thiazoles with a Ph substituent at C(2), the reaction proceeded via formal [2+2] cycloaddition and led to the correspoding β‐lactam. On the other hand, acid chlorides and 4,5‐dihydro‐1,3‐thiazoles bearing an α‐H‐atom at the C(2)‐substituent underwent C(α)‐ and/or N‐addition reactions and furnished non‐β‐lactam adducts, i.e., C(α)‐ and/or N‐acylated 1,3‐thiazolidines. The attempted transformations of sulfonyl esters of exo‐6‐hydroxy penams to endo‐6‐azido penams failed, although they were successful with mono‐β‐lactams under the same conditions.  相似文献   

16.
Three new lycopodium alkaloids, huperserramines A–C ( 1 – 3 , resp.), along with 15 known ones, lycopodine‐6α,11α‐diol ( 4 ), lycoposerramine H ( 5 ), lycoposerramine I ( 6 ), lycopodine‐6α‐ol ( 7 ), lycoposerramine M ( 8 ), diphaladine A ( 9 ), lycoposerramine K ( 10 ), lycoposerramine W ( 11 ), huperzine M ( 12 ), luciduline ( 13 ), phlegmariuine N ( 14 ), huperzine A ( 15 ), huperzine B ( 16 ), lycodine ( 17 ), and lycoposerramine R ( 18 ), were isolated from the whole plant of Huperzia serrata. Their structures were established by spectroscopic methods, including 2D‐NMR and MS analyses. All the isolates were evaluated for their inhibitory effects on acetylcholinesterase (AChE) and α‐glucosidase. As a result, lycopodine‐6α,11α‐diol ( 4 ) exhibited more potent α‐glucosidase inhibitory activity (IC50 148±5.5 μM ) than the positive control acarbose (IC50 376.3±2.7 μM ).  相似文献   

17.
Racemic and enantiomerically pure manno‐configured isoquinuclidines were synthesized and tested as glycosidase inhibitors. The racemic key isoquinuclidine intermediate was prepared in high yield by a cycloaddition (tandem Michael addition/aldolisation) of the 3‐hydroxy‐1‐tosyl‐pyridone 10 to methyl acrylate, and transformed to the racemic N‐benzyl manno‐isoquinuclidine 2 and the N‐unsubstituted manno‐isoquinuclidine 3 (twelve steps; ca. 11% from 10 ). Catalysis by quinine of the analogous cycloaddition of 10 to (?)‐8‐phenylmenthyl acrylate provided a single diastereoisomer in high yield, which was transformed to the desired enantiomerically pure D ‐manno‐isoquinuclidines (+)‐ 2 and (+)‐ 3 (twelve steps; 23% from 10 ). The enantiomers (?)‐ 2 and (?)‐ 3 were prepared by using a quinidine‐promoted cycloaddition of 10 to the enantiomeric (+)‐8‐phenylmenthyl acrylate. The N‐benzyl D ‐manno‐isoquinuclidine (+)‐ 2 is a selective and slow inhibitor of snail β‐mannosidase. Its inhibition strength and type depends on the pH (at pH 4.5: Ki=1.0 μM , mixed type, α=1.9; at pH 5.5: Ki=0.63 μM , mixed type, α=17). The N‐unsubstituted D ‐manno‐isoquinuclidine (+)‐ 3 is a poor inhibitor. Its inhibition strength and type also depend on the pH (at pH 4.5: Ki=1.2?103 μM , mixed type, α=1.1; at pH 5.5: Ki=0.25?103 μM , mixed type, α=11). The enantiomeric N‐benzyl L ‐manno‐isoquinuclidine (?)‐ 2 is a good inhibitor of snail β‐mannosidase, albeit noncompetitive (at pH 4.5: Ki=69 μM ). The N‐unsubstituted isoquinuclidine (?)‐ 2 is a poor inhibitor (at pH 4.5: IC50=7.3?103 μM ). A comparison of the inhibition by the pure manno‐isoquinuclidines (+)‐ 2 and (+)‐ 3 , (+)‐ 2 /(?)‐ 2 1 : 1, and (+)‐ 3 /(?)‐ 3 1 : 1 with the published data for racemic 2 and 3 led to a rectification of the published data. The inhibition of snail β‐mannosidase by the isoquinuclidines 2 and 3 suggests that the hydrolysis of β‐D ‐mannopyranosides by snail β‐mannosidase proceeds via a distorted conformer, in agreement with the principle of stereoelectronic control.  相似文献   

18.
Four new podocarpane‐type trinorditerpenenes, (5β,10α)‐12,13‐dihydroxypodocarpa‐8,11,13‐trien‐3‐one ( 1 ), (5β,10α)‐12‐hydroxy‐13‐methoxypodocarpa‐8,11,13‐trien‐3‐one ( 2 ), (5β,10α)‐13‐hydroxy‐12‐methoxypodocarpa‐8,11,13‐trien‐3‐one ( 3 ), and (3α,5β,10α)‐13‐methoxypodocarpa‐8,11,13‐triene‐3,12‐diol ( 4 ), together with four known diterpenes, 12‐hydroxy‐13‐methylpodocarpa‐8,11,13‐trien‐3‐one ( 5 ), spruceanol ( 6 ), ent‐3α‐hydroxypimara‐8(14),15‐dien‐12‐one ( 7 ), and ent‐3β,14α‐hydroxypimara‐7,9(11),15‐triene‐12‐one ( 8 ), were isolated from the twigs and leaves of Aleurites moluccana. Their structures were elucidated by means of comprehensive spectroscopic analyses, including NMR and MS. Except 8 , all compounds were evaluated for their cytotoxicity; compound 4 exhibited moderate inhibitory activity against Raji cells with an IC50 value of 4.24 μg/ml.  相似文献   

19.
Four new ent‐pimarane diterpenes were isolated from the EtOH extract of Aralia dumetorum, together with three known compounds involving ent‐pimar‐8(14),15‐dien‐19‐oic acid ( 5 ), ent‐pimar‐8(14),15‐dien‐19‐ol ( 6 ), and ent‐kaur‐16‐en‐19‐oic acid ( 7 ). By detailed analyses of the MS, IR, and NMR data, the structures of four new diterpenes were characterized as (5β,9β,10α,13α)‐pimara‐6,8(14),15‐trien‐18‐oic acid ( 1 ), (5β,7β,9β,10α,13α)‐7‐methoxypimara‐8(14),15‐dien‐18‐oic acid ( 2 ), (5β,9β,10α,13α,14β)‐14‐methoxypimara‐7,15‐dien‐18‐oic acid ( 3 ), and (5β,10α,13α,14α)‐14‐hydroxypimara‐7,9(11),15‐trien‐18‐oic acid ( 4 ). The cytotoxic activities of compounds 1  –  7 were assayed in vitro through MTT method.  相似文献   

20.
In the present study, it was demonstrated that ethyl acetate soluble fraction partitioned from heartwood of Dalbergia odorifera T. Chen (HEF) had a remarkable inhibitory effect on α‐glucosidase. Therefore HEF was selected as a starting material for screening the potential α‐glucosidase inhibitors using ultrafiltration liquid chromatography/mass spectrometry (UF‐LC/MS). Twenty‐six compounds were identified with analysis of LC/MS. UF assay indicated that 18 compositions might be α‐glucosidase inhibitors in HEF; eight of them were estimated for their α‐glucosidase inhibitory activity, and the results showed that (2S)‐liquiritigenin, (2S)‐4′,6‐dihydroxy‐ 7‐methoxyflavanone and isoliquiritigenin displayed obvious inhibition of yeast α‐glucosidase. In addition, in order to control the quality of HEF, the content of five compounds in HEF was simultaneously determined for the first time. These results provide an important theoretical base for the further application of HEF to treat type 2 diabetes in the clinic and development of natural α‐glucosidase inhibitors with low toxicity. Copyright © 2013 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