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 55 毫秒
1.
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   

2.
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   

3.
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   

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.
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   

6.
The (−)‐ and (+)‐β‐irones ((−)‐ and (+)‐ 2 , resp.), contaminated with ca. 7 – 9% of the (+)‐ and (−)‐transα‐isomer, respectively, were obtained from racemic α‐irone via the 2,6‐trans‐epoxide (±)‐ 4 (Scheme 2). Relevant steps in the sequence were the LiAlH4 reduction of the latter, to provide the diastereoisomeric‐4,5‐dihydro‐5‐hydroxy‐transα‐irols (±)‐ 6 and (±)‐ 7 , resolved into the enantiomers by lipase‐PS‐mediated acetylation with vinyl acetate. The enantiomerically pure allylic acetate esters (+)‐ and (−)‐ 8 and (+)‐ and (−)‐ 9 , upon treatment with POCl3/pyridine, were converted to the β‐irol acetate derivatives (+)‐ and (−)‐ 10 , and (+)‐ and (−)‐ 11 , respectively, eventually providing the desired ketones (+)‐ and (−)‐ 2 by base hydrolysis and MnO2 oxidation. The 2,6‐cis‐epoxide (±)‐ 5 provided the 4,5‐dihydro‐4‐hydroxy‐cisα‐irols (±)‐ 13 and (±)‐ 14 in a 3 : 1 mixture with the isomeric 5‐hydroxy derivatives (±)‐ 15 and (±)‐ 16 on hydride treatment (Scheme 1). The POCl3/pyridine treatment of the enantiomerically pure allylic acetate esters, obtained by enzymic resolution of (±)‐ 13 and (±)‐ 14 , provided enantiomerically pure cisα‐irol acetate esters, from which ketones (+)‐ and (−)‐ 22 were prepared (Scheme 4). The same materials were obtained from the (9S) alcohols (+)‐ 13 and (−)‐ 14 , treated first with MnO2, then with POCl3/pyridine (Scheme 4). Conversely, the dehydration with POCl3/pyridine of the enantiomerically pure 2,6‐cis‐5‐hydroxy derivatives obtained from (±)‐ 15 and (±)‐ 16 gave rise to a mixture in which the γ‐irol acetates 25a and 25b and 26a and 26b prevailed over the α‐ and β‐isomers (Scheme 5). The (+)‐ and (−)‐cisγ‐irones ((+)‐ and (−)‐ 3 , resp.) were obtained from the latter mixture by a sequence involving as the key step the photochemical isomerization of the α‐double bond to the γ‐double bond. External panel olfactory evaluation assigned to (+)‐β‐irone ((+)‐ 2 ) and to (−)‐cisγ‐irone ((−)‐ 3 ) the strongest character and the possibility to be used as dry‐down note.  相似文献   

7.
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   

8.
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   

9.
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   

10.
New 2‐(aminomethyl)‐5‐(hydroxymethyl)pyrrolidine‐3,4‐diol derivatives were synthesized from (5S)‐5‐[(trityloxy)methyl]pyrrolidin‐2‐one ( 6 ) (Schemes 1 and 2) and their inhibitory activities toward 25 glycosidases assayed (Table). The influence of the configuration of the pyrrolidine ring on glycosidase inhibition was evaluated. (2R,3R,4S,5R)‐2‐[(benzylamino)methyl]‐5‐(hydroxymethyl)pyrrolidine‐3,4‐diol ((+)‐ 21 ) was found to be a good and selective inhibitor of α‐mannosidase from jack bean (Ki=1.2 μM ) and from almond (Ki=1.0 μM ). Selectivity was lost for the non‐benzylated derivative (2R,3R,4S,5R)‐2‐(aminomethyl)‐5‐(hydroxymethyl)pyrrolidine‐3,4‐diol ((+)‐ 22 ) which inhibited α‐galactosidases, β‐galactosidases, β‐glucosidases, and αN‐acetylgalactosaminidase as well.  相似文献   

11.
The cross‐aldolization of (−)‐(1S,4R,5R,6R)‐6‐endo‐chloro‐5‐exo‐(phenylseleno)‐7‐oxabicyclo[2.2.1]heptan‐2‐one ((−)‐ 25 ) and of (+)‐(3aR,4aR,7aR,7bS)‐ ((+)‐ 26 ) and (−)‐(3aS,4aS,7aS,7bR)‐3a,4a,7a,7b‐tetrahydro‐6,6‐dimethyl[1,3]dioxolo[4,5]furo[2,3‐d]isoxazole‐3‐carbaldehyde ((−)‐ 26 ) was studied for the lithium enolate of (−)‐ 25 and for its trimethylsilyl ether (−)‐ 31 under Mukaiyama's conditions (Scheme 2). Protocols were found for highly diastereoselective condensation giving the four possible aldols (+)‐ 27 (`anti'), (+)‐ 28 (`syn'), 29 (`anti'), and (−)‐ 30 (`syn') resulting from the exclusive exo‐face reaction of the bicyclic lithium enolate of (−)‐ 25 and bicyclic silyl ether (−)‐ 31 . Steric factors can explain the selectivities observed. Aldols (+)‐ 27 , (+)‐ 28 , 29 , and (−)‐ 30 were converted stereoselectively to (+)‐1,4‐anhydro‐3‐{(S)‐[(tert‐butyl)dimethylsilyloxy][(3aR,4aR,7aR,7bS)‐3a,4a,7a,7b‐tetrahydro‐6,6‐dimethyl[1,3]dioxolo[4,5]‐furo[2,3‐d]isoxazol‐3‐yl]methyl}‐3‐deoxy‐2,6‐di‐O‐(methoxymethyl)‐α‐D ‐galactopyranose ((+)‐ 62 ), its epimer at the exocyclic position (+)‐ 70 , (−)‐1,4‐anhydro‐3‐{(S)‐[(tert‐butyl)dimethylsilyloxy][(3aS,4aS,7aS,7bR)‐3a,4a,7a,7b‐tetrahydro‐6,6‐dimethyl[1,3]dioxolo[4,5]furo[2,3‐d]isoxazol‐3‐yl]methyl}‐3‐deoxy‐2,6‐di‐O‐(methoxymethyl)‐α‐D ‐galactopyranose ((−)‐ 77 ), and its epimer at the exocyclic position (+)‐ 84 , respectively (Schemes 3 and 5). Compounds (+)‐ 62 , (−)‐ 77 , and (+)‐ 84 were transformed to (1R,2R,3S,7R,8S,9S,9aS)‐1,3,4,6,7,8,9,9a‐octahydro‐8‐[(1R,2R)‐1,2,3‐trihydroxypropyl]‐2H‐quinolizine‐1,2,3,7,9‐pentol ( 21 ), its (1S,2S,3R,7R,8S,9S,9aR) stereoisomer (−)‐ 22 , and to its (1S,2S,3R,7R,8S,9R,9aR) stereoisomer (+)‐ 23 , respectively (Schemes 6 and 7). The polyhydroxylated quinolizidines (−)‐ 22 and (+)‐ 23 adopt `trans‐azadecalin' structures with chair/chair conformations in which H−C(9a) occupies an axial position anti‐periplanar to the amine lone electron pair. Quinolizidines 21 , (−)‐ 22 , and (+)‐ 23 were tested for their inhibitory activities toward 25 commercially available glycohydrolases. Compound 21 is a weak inhibitor of β‐galactosidase from jack bean, of amyloglucosidase from Aspergillus niger, and of β‐glucosidase from Caldocellum saccharolyticum. Stereoisomers (−)‐ 22 and (+)‐ 23 are weak but more selective inhibitors of β‐galactosidase from jack bean.  相似文献   

12.
The racemic 2‐azabicyclo[3.2.2]nonanes 5 and 18 were synthesized and tested as β‐glycosidase inhibitors. The intramolecular Diels–Alder reaction of the masked o‐benzoquinone generated from 2‐(allyloxy)phenol ( 6 ) gave the α‐keto acetal 7 which was reduced with SmI2 to the hydroxy ketone 8 . Dihydroxylation, isopropylidenation (→ 12 ), and Beckmann rearrangement provided lactam 15 . N‐Benzylation of this lactam, reduction to the amine 17 , and deprotection provided the amino triol 19 which was debenzylated to the secondary amine 5 . Both 5 and 19 proved weak inhibitors of snail β‐mannosidase (IC50 > 10 mM ), Caldocellum saccharolyticum β‐glucosidase (IC50 > 10 mM ), sweet almond β‐glucosidase (IC50 > 10 mM ), yeast α‐glucosidase ( 5 : IC50 > 10 mM ; 19 : IC50 = 1.2 mM ), and Jack bean α‐mannosidase (no inhibition detected).  相似文献   

13.
Two independent total syntheses of the Aristotelia alkaloid (−)‐serratenone ((−)‐ 1 ) are disclosed, one starting with (−)‐α‐pinene, the other one with (S)‐α‐terpineol. These correlations led to a revision of the originally proposed absolute configuration of the natural product. In the course of systematic investigations of the behavior of the indole alkaloids (+)‐makomakine ((+)‐ 18 ) and (−)‐hobartine ((−)‐ 22 ) towards oxidizing reagents, it was found that treatment with I2 leads to no less than five different products. Depending on the exact reaction conditions, each of them can be obtained as the major component in yields between 40 and 60%. One of these compounds was shown to be identical with the natural product (+)‐11,12‐didehydromakonin‐10‐one ((+)‐ 28 ).  相似文献   

14.
To complete our panorama in structure–activity relationships (SARs) of sandalwood‐like alcohols derived from analogues of α‐campholenal (= (1R)‐2,2,3‐trimethylcyclopent‐3‐ene‐1‐acetaldehyde), we isomerized the epoxy‐isopropyl‐apopinene (?)‐ 2d to the corresponding unreported α‐campholenal analogue (+)‐ 4d (Scheme 1). Derived from the known 3‐demethyl‐α‐campholenal (+)‐ 4a , we prepared the saturated analogue (+)‐ 5a by hydrogenation, while the heterocyclic aldehyde (+)‐ 5b was obtained via a Bayer‐Villiger reaction from the known methyl ketone (+)‐ 6 . Oxidative hydroboration of the known α‐campholenal acetal (?)‐ 8b allowed, after subsequent oxidation of alcohol (+)‐ 9b to ketone (+)‐ 10 , and appropriate alkyl Grignard reaction, access to the 3,4‐disubstituted analogues (+)‐ 4f,g following dehydration and deprotection. (Scheme 2). Epoxidation of either (+)‐ 4b or its methyl ketone (+)‐ 4h , afforded stereoselectively the trans‐epoxy derivatives 11a,b , while the minor cis‐stereoisomer (+)‐ 12a was isolated by chromatography (trans/cis of the epoxy moiety relative to the C2 or C3 side chain). Alternatively, the corresponding trans‐epoxy alcohol or acetate 13a,b was obtained either by reduction/esterification from trans‐epoxy aldehyde (+)‐ 11a or by stereoselective epoxidation of the α‐campholenol (+)‐ 15a or of its acetate (?)‐ 15b , respectively. Their cis‐analogues were prepared starting from (+)‐ 12a . Either (+)‐ 4h or (?)‐ 11b , was submitted to a Bayer‐Villiger oxidation to afford acetate (?)‐ 16a . Since isomerizations of (?)‐ 16 lead preferentially to β‐campholene isomers, we followed a known procedure for the isomerization of (?)‐epoxyverbenone (?)‐ 2e to the norcampholenal analogue (+)‐ 19a . Reduction and subsequent protection afforded the silyl ether (?)‐ 19c , which was stereoselectively hydroborated under oxidative condition to afford the secondary alcohol (+)‐ 20c . Further oxidation and epimerization furnished the trans‐ketone (?)‐ 17a , a known intermediate of either (+)‐β‐necrodol (= (+)‐(1S,3S)‐2,2,3‐trimethyl‐4‐methylenecyclopentanemethanol; 17c ) or (+)‐(Z)‐lancifolol (= (1S,3R,4Z)‐2,2,3‐trimethyl‐4‐(4‐methylpent‐3‐enylidene)cyclopentanemethanol). Finally, hydrogenation of (+)‐ 4b gave the saturated cis‐aldehyde (+)‐ 21 , readily reduced to its corresponding alcohol (+)‐ 22a . Similarly, hydrogenation of β‐campholenol (= 2,3,3‐trimethylcyclopent‐1‐ene‐1‐ethanol) gave access via the cis‐alcohol rac‐ 23a , to the cis‐aldehyde rac‐ 24 .  相似文献   

15.
N‐Substituted (3S,4S)‐ and (3R,4R)‐pyrrolidine‐3,4‐diols 9 and 10 , respectively, were derived from (+)‐L ‐ and (?)‐D ‐tartaric acid, respectively. Compounds 9k, 9l , and 9m with the N‐substituents, BnNH(CH2)2, 4‐PhC6H4CH2NH(CH2)2 and 4‐ClC6H4CH2NH(CH2)2, respectively, showed modest inhibitory activities toward α‐D ‐amyloglucosidases from Aspergillus niger and from Rhizopus mold (Table 1). Unexpectedly, several (3R,4R)‐pyrrolidine‐3,4‐diols 10 showed inhibitory activities toward α‐D ‐mannosidases from almonds and from jack bean (Table 3). N‐Substitution by the NH2(CH2)2 group, i.e., 10g , led to the highest potency.  相似文献   

16.
The inhibition of the β‐glucosidases from sweet almonds and from Caldocellum saccharolyticum by the 4‐amino‐4‐deoxy lactam 11 , the 4‐deoxy lactam 12 , and the corresponding imidazoles 13 and 14 was compared to the inhibition by the hydroxy analogues 1 and 3 . Substitution of the OH group at C(4) by an amino group or by hydrogen weakened the inhibition by ΔΔGdiss = + 1.9 to + 3.1 kcal/mol. Similarly, the inhibition of the β‐galactosidase from bovine liver and from E. coli by the 4‐deoxy lactam 12 and the imidazole 14 , as compared to the one by the galacto‐configured lactam 9 and imidazole 10 , is weakened by deoxygenation at C(4) (ΔΔGdiss = + 2.6 and 4.5 kcal/mol, resp.). The effect of these substitutions on the inhibition of the C. saccharolyticum β‐glucosidase is slightly stronger than the one on the sweet almonds β‐glucosidases. The effect is also stronger on the inhibition by the imidazoles than by the lactams, and depends on the flexibility of the inhibitors. The amino and deoxy lactams 11 and 12 were prepared from the galactonolactam‐derived triflate 17 by substitution with azide and hydride, respectively, followed by hydrogenation. Azidation of the galacto‐configured imidazopyridine‐derived triflate 24 and hydrogenation gave the amino‐imidazole 13 . The deoxy lactam 20 was transformed to the manno‐ and gluco‐configured deoxy‐imidazoles 29 and 30 via the thionolactam 28 . Hydrogenolytic deprotection of 30 gave the deoxy‐imidazole 14 .  相似文献   

17.
The C(2)‐substituted imidazoles 11 , 15 – 17 , 19 , 21 , 23 / 24 , 28 – 31 , 37 , and 38 have been prepared from the known 2,3‐unsubstituted imidazole 7 via the iodoimidazole 10 , and tested as inhibitors of β‐ and α‐glucosidases. Introduction of hydrophobic and flexible substituents, such as in 28 and 29 , led to a very strong inhibition of β‐glucosidases, with Ki values for 29 of 1.2 and 0.11 nM against β‐glucosidases from almonds and Caldocellum saccharolyticum, respectively. A slow onset of the inhibition was observed for the strongly inhibiting 16 , 28 – 31 , 37 , and 38 . While the introduction of a hydroxymethyl or a phenethyl substituent as in 17 and 30 led to stronger inhibition, the 1′‐hydroxyphenethyl derivatives 37 and 38 were weaker inhibitors than 16 and 29 . This result is interpreted in the light of a conformational change of the substrate on the way to the transition state. The substituent at C(2) has only a moderate influence on the selectivity of the inhibition of two β‐ and one α‐glucosidases, increasing it by a maximal factor of ca. 10 ( 16 ), or decreasing it by a maximal factor of ca. 15 ( 37 ).  相似文献   

18.
The cyclic [2R,S(R)]‐bornane‐10,2‐sulfinamide (−)‐ 2b , an analogue of Oppolzer`s camphor‐derived sultam (−)‐ 2a , was synthesized by reduction of the known N‐alkylidenesulfinamide (+)‐ 1b with NaBH4. The uncatalyzed [4+2] cycloaddition of cyclopentadiene to the methyl ester (−)‐ 3b of the N‐fumaroylsulfinamide, obtained from (−)‐ 2b , proceeds with lower endo and π‐facial selectivity as compared to dienophiles (−)‐ 3a , c . In contrast to these latter, the diastereoselectivity is reversed either in apolar CCl4 or in the presence of TiCl4. This inversion is explained by a competitive C(α)‐si addition on the reactive anti‐s‐trans conformer.  相似文献   

19.
In this work, peripheral or nonperipheral tetra‐[4‐(9H‐carbazol‐9‐yl)phenoxy] substituted cobalt(II), manganese (III) phthalocyanines were synthesized for the first time. Their acetylcholinesterase from Electrophorus electricus (AChE), butyrylcholinesterase equine serum (BuChE), and α‐glucosidase Saccharomyces cerevisiae inhibition were investigated spectrophotometrically. Finally, in vitro cytotoxicities of the compounds were investigated on human neuroblastoma (SH‐SY5Y) cell line using MTT cell viability assay. The compounds inhibited to enzymes in the range of 7.39 ± 0.25–35.29 ± 2.49 μM with IC50 values for AChE and 14.38 ± 0.66–58.02 ± 4.94 μM for BuChE as compared with galantamine, which used as a positive control. For α‐glucosidase, all compounds had stronger inhibition action than acarbose according to the IC50 values. The IC50 values of N? Co and N? Mn were found to be 3.05 ± 0.10 and 15.82 ± 1.85 μM, respectively. The results of cytotoxicity showed that the IC50 values were above 100 μM showing the compounds had low cytotoxic action against SH‐SY5Y cell line for 24 h. Overall, carbazole substituted nonperipheral compounds can be considered as a potential agent for the treatment of Alzheimer's diseases and diabetes mellitus.  相似文献   

20.
The four hederagenin glycosides 1 – 4 were efficiently synthesized through one‐pot sequential glycosylations with glycose 1‐(trichloroacetimidate)s as donors, resulting in a significantly simplified synthetic procedure without isolation of glycosylation intermediates. The activity of the synthetic hederagenin glycosides 1 – 4 against α‐glucosidase type IV was evaluated; hederagenin glycoside 4 containing an α‐L ‐rhamnopyranosyl unit showed the best activity with an IC50 value of 47.9 μM .  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