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 46 毫秒
1.
At pH 4.5 (citrate buffer), D -gluconhydroximo-lactone ( 2 ), the N-methylurethane 3 and the N-phenylurethane 4 inhibit competitively the hydrolysis of p-nitrophenyl β-D -glucopyranoside by emulsin. The IC50 values of 2, 3 , and 4 were 1.6 × 10?4, 1.0 × 10?4, and 5.8 × 10?6 M , respectively. The Ki values of 2 and 4 were 9.8 × 10?5 and 2.3 × 10?6 M , respectively, while D-glucono-1,5-lactone ( 1 ) showed IC50 = 1.1 × 10?4 M and Ki = 3.7 × 10?5 M .  相似文献   

2.
The (1H-imidazol-2-yl)ulose 8 and the 1H-imidazol-2-yl C-glycopyranosides 23 and 24 have been prepared from tetra-O-benzylgluconolactone 6 in two and six steps, respectively. The imidazoles 8 and 24 are moderate competitive inhibitors of sweet-almond β-glucosidase (pH 6.8, Ki ≈ 0.79 and 0.64 mM , respectively), while 23 is a competitive inhibitor of yeast α-glucosidase (pH 6.8, Ki ≈ 0.26 mM ). Addition of 2-lithiated 1-[(dimethylamino)methyl]-1H-imidazole to 6 gave the ulose 7 (68%), which was deprotected to 8 . Reduction of 7 with NaBH4 yielded a 12:88 mixture 10/11 . Attempts to selectively mesylate HO? C(1) of these diols failed, while dinitrobenzoylation led to 19/20 , which cyclized easily (NaH) to a 25:75 mixture of 21 and 22 which were separated and debenzylated to the C-glycosides 23 and 24 .  相似文献   

3.
Epoxy type inhibitors, 3‐t‐butylphenyl 3‐1,2‐epoxybutyl ether ( 1 ), 3‐t‐butylphenyl 3‐1,2‐epoxyhexyl ether ( 2 ), and 2‐naphthyl 3‐1,2‐epoxyhexyl ether ( 3 ) are synthesized as the active site‐directed inhibitors of cholesterol esterase, acetylcholinesterase, and butyrylcholinesterase. All epoxy compounds are characterized as the time‐independent inhibitors for all three enzymes from the stopped‐time assay. Further, all epoxy compounds are characterized as the competitive inhibitors for all three enzymes from the Lineweaver‐Burk plots. The inhibition constants (Ki) of cholesterol esterase for compounds 1‐3 are 320 ± 40, 190 ± 20, 130 ± 20 μM, respectively. The Ki values of acetylcholinesterase for compounds 1‐3 are 490 ± 20, 141 ± 5, 200 ± 30 μM, respectively. Values of Ki of butyrylcholinesterase for compounds 1‐3 are 250 ± 30, 26 ± 4, 120 ± 20 μM, respectively. Compound 2 is the most potent inhibitor for butyrylcholinesterase probably because the compound mimics most the natural substrate, butyrylcholine.  相似文献   

4.
The phosphonic acids 3 and 4 were prepared to compare their inhibitory activity on Vibrio cholerae sialidase with the one of the corresponding N-acetyl-2-deoxyneuraminic acids 5 and 6 . Thus, hydrogenation and benzylation of methyl N-acetyl-2,3-didehydro-2-deoxyneuraminate (1MeNeu2en5Ac; 7) gave a mixture of the fully O-benzylated benzyl and methyl esters 9 and 10 , the partially O-benzylated benzyl and methyl esters 11 and 12 , and the fully O-and N-benzylated benzyl and methyl esters 13 and 14 (Scheme 1). Transesterification of 9 to 10 and hydrolysis of 10 gave the acid 15 . Oxidative decarboxylation of 15 with Pb(OAc)4 gave a 1:9 mixture of the α-and β-D-glycero-D-galacto-acetates 16 and 17 . Phosphonoylation of 17 with P(OMe)3 and Me3SiOTf gave a 1.3:1 mixture of the phosphonates 18 and 19 , which were deprotected to give the (4-acetamido-2,4-dideoxy-D-glycero-α-and β-D-galacto-octopyranosyl)phosphonic acids 3 and 4 , respectively. The acid 6 was obtained by epimerization of the tert-butyl ester 23 with lithium N-cyclohexylisoproylamide and deprotection. The phosphonic acids 3 (Ki 5.5 10-5 M) and 4 (Ki 2.3.10?4 M ) are stronger inhibitors of Vibrio cholerae sialidase than the anomeric N-acetyl-2-deoxyneuraminic acids 5 (Ki 2.3 10?3 M ) and 6 . Both 3 and 4 inhibit the Vibrio cholerae sialidase, while only the carboxylic acid 5 , possessing an equatorial COOH group is an inhibitor.  相似文献   

5.
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   

6.
Herein, novel silver-based N-heterocyclic carbene (NHC) complexes bearing 2-(4-hydroxyphenyl)ethyl group were synthesized. Novel Ag(I)NHC complexes were synthesized from the 2-(4-hydroxyphenyl)ethyl-substituted benzimidazolium salts and silver oxide via in situ deprotonation method. The successful formation of all Ag(I)NHC complexes was proved by using 1H NMR, 13C NMR, FTIR spectroscopy, and elemental analysis techniques. In addition, their inhibitory effects have been investigated of these substances on acetylcholinesterase (AChE), α-glycosidase (α-Gly), human carbonic anhydrase I (hCA I), and human carbonic anhydrase II (hCA II) enzymes. It has been seen that all compounds have a better ability to inhibit compared with existing tried inhibitors. Among these, the best inhibitor against AChE enzyme is 1g (Ki : 9.54 ± 0.98 μM and IC50 : 17.40), and against α-Gly, 1c showed the highest effect (Ki 3.09 ± 0.36 μM and IC50 7.91). The best inhibitor against hCA I and hCA II enzymes are 1c and 1g compounds. For hCA I and hCA II, IC50 values were calculated as 17.85 and 9.06 μM and Ki values were measured as 5.45 ± 2.02 and 8.99 ± 2.02 μM, respectively.  相似文献   

7.
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   

8.
The racemic 7‐oxanorbornanyl and norbornanyl aminoalcohols 3, 4, 42, 45 , and 46 were synthesized and tested as snail β‐mannosidase inhibitors. The amino tetraol 3 was obtained from the known sulfonyl acrylate 9 and furan 10 . Esterification provided 11 that underwent an intramolecular Diels–Alder reaction to the 7‐oxanorbornene 12 . Reduction of 12 to 13 , desulfonylation, isopropylidenation, and cis‐dihydroxylation gave 16 . A second isopropylidenation to 17 , followed by debenzylation and a Mitsunobu–Gabriel reaction provided 19 that was deprotected via 20 to 3 . Diels–Alder cycloaddition of furfuryl acetate and maleic anhydride to 21 , followed by alcoholysis of the anhydride, cis‐dihydroxylation, isopropylidenation, and Barton decarboxylation gave the ester 25 . Deacetylation to 26 and a Mitsunobu–Gabriel reaction led to 27 that was transformed into the N‐Boc analogue 29 , reduced to the alcohol 30 , and deprotected to 4 . The 1‐aminonorbornane 5 was obtained from Thiele's Acid 31 . Diels–Alder cycloaddition of the cyclopentadiene obtained by thermolysis of the diester 32 , methanolysis of the resulting anhydride 33 , dihydroxylation, isopropylidenation, Barton decarboxylation, and Curtius degradation led to the benzyl carbamate 39 that was reduced to the alcohol 40 , transformed into the N‐Boc carbamate 41 , and deprotected to 5 . The alcohol 40 was also transformed into the benzylamine 42 , aniline 45 , and hydroxylamine 46 . Snail β‐mannosidase was hardly inhibited by 3, 4, 42, 45 , and 46 . Only the amino triol 5 proved a stronger inhibitor. The inhibition by 5 depends on the pH value (at pH 3.5: Ki = 1900 μM ; at pH 4.5: Ki = 340 μm; at pH 5.5: Ki = 110 μm). The results illustrate the strong dependence of the inhibition by bicyclic mimics upon the precise geometry and orientation of the amino group as determined by the scaffold. It is in keeping with the hypothesis that the reactive conformation imposed by snail β‐mannosidase is close to a 1,4B/1S3.  相似文献   

9.
The N‐acetylglucosamine‐related 1,2,4‐triazole 14 and 1,2,3‐triazole 16 have been prepared by N‐acetylation of the known amines 19 and 20 , and their Ki values determined against bovine kidney βN‐acetylglucosaminidase, a mammalian hexosaminidase. The 1,2,3‐triazole 16 (Ki=4 μM ) is a markedly weaker inhibitor than the isosteric azoles 13 – 15 . The Ki value of the 1,2,4‐triazole 14 (0.034 μM ) is smaller than that of the tetrazole 13 (0.2 μM ), but larger than that of the imidazole 15 (0.0035 μM ), confirming the correlation between inhibitory strength and basicity of the azole, as expected on the basis of an anti‐protonation mechanism of mammalian hexosaminidases.  相似文献   

10.
Nα-[4-[[(4-Aminopteridin-6-yl)methyl]amino]benzoyl]-L-ornithine (dAPA-Orn) was synthesized, and its ability to inhibit folylpolyglutamate synthetase from mouse liver was compared with that of the corresponding 2,4-diamino analogue APA-Orn. Also compared were the inhibitory activities of the deaza analogues 5-deazaAPA-Orn, 8-deazaAPA-Orn, and 5,8-dideazaAPA-Orn, as well as those of Nα-pteroyl-L-ornithine (PteOrn) and its deaza analogues 5-deazaPteOrn and 5,8-dideazaPteOrn. The inhibition constant Ki of dAPA-Orn was 7-fold greater than that of APA-Orn, indicating that the 2-amino group plays a role in binding to the active site. The binding affinity of the 2,4-diamino compounds increased in the order 5-deazaAPA < APA-Orn <5,8-dideazaAPA-Orn < 8-deazaAPA-Orn, and that of the 2-amino-4(3H)-oxo compounds increased in the order 5-deazaPteOrn < PteOrn < 5,8-dideazaPteOrn. The most potent inhibitor of both groups was 8-deazaAPA-Orn, with a Ki of 0.018 μM, coresponding to an 8-fold and 15-fold increase in affinity relative to APA-Orn and 5-deazaAPA-Orn, respectively. The results suggest (a) that the binding of Orn-containing folylpolyglutamate synthetase inhibitors is affected to a greater degree by replacement of N8 by a carbon atom than it is by the corresponding change at N5, (b) that the effect of carbon for nitrogen replacement is greater in the 2,4-diamino derivatives than in the 2-amino-4(3H)-oxo compounds, and (c) that the 2,4-diamines are the better inhibitors. Comparison of the Ki values of the Orn-containing inhibitors with the Km values of the corresponding glutamate-containing substrates revealed that Km/Ki ratio can vary as much as 100-fold depending on the nature of the heterocyclic moiety, suggesting that caution should be exercised in using Km values of known substrates to predict Ki values of putative inhibitors.  相似文献   

11.
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   

12.
(−)‐ 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   

13.
The thiolactone oxime 10 was synthesized in ten steps from the known tri-O-benzylglucose 13 , which was transformed into the oxime 14 , silylated (→ 15 ), and mesylated (→ 16 ). Treatment of 16 with Bu4NF yielded the L -ido-epoxide 17 and the hydroxylamine 18 ; the isomeric D -gluco-configurated hydroxylamine 20 was prepared from 17 . Reaction of 17 with thiourea yielded the thiirane 19 . Ring opening was best effected with HBr (→ 22 ·HBr). The N-glycosylhydroxylamine 22 was immediately oxidized to 24 , as it reverted to 19 . Similarly, 19 was transformed into the chlorides 21 and 23 . The iodide 25 reacted with TEMPO to afford 29 besides 26 and 30 ; nucleophilic substitution of 23 , 24 , or 25 gave unsatisfactory yields of 26 or 27 , and 28 . Birch reduction transformed 29 into 10 which was isolated via the pentaacetate 32 , which was also transformed into the tetraacetate 33 . The weak activity of 10 as an inhibitor of sweet-almond and Agrobacter β-glucosidase is in keeping with categorization of the lactone and lactam oximes 1–5 and the 5-thiosugars 6–9 as transition-state and substrate analogs, respectively.  相似文献   

14.
A sweet almond β-glucosidase (EC 3.2.1.21) isozyme was purified from commercial crude product. The process of purification consisted of a Protein-Pak Q anion exchange chromatography following by a Superdex 75 HR gel filtration separation. The purified enzyme is a monomeric glycoprotein with molecular weight of 58 kDa and pI=4.55 which is distinguished from reported isozymes. The enzyme has apH optimum in the range of 5.2-5.6 when p-nitrophenyl-β-D-glycopyranosides are used as substrate and is stable up to 50 °C at that pH range. The purified protein also exhibits profound β-galactosidase and σ-L-arabinosidase activity. The study of substrate specificity revealed that lacking of hydroxymethyl group at C-5 of glycosides resulted in higher affinity for substrate binding to enzyme, whereas the chemical step of hydrolysis (kcst) was prevented significantly. The pH activity profile displayed a bell-shaped curve for all measured p-nitrophenyl-β-D-glycopyranosides with apparent pK1 and pK2 values of 4.4-4.7 and 6.2-6.4, respectively. This isozyme was strongly inhibited by δ-gluconolactone (Ki = 160 μM) and 4-phenylimidazole (Ki = 17.8 μM) reversibly at pH 6.2. Among the tested glycoses, the binding affinity of N-acetyl-β-D-glucosamine to the enzyme (Kl = 52 mM) was 6 times stronger than that of glucose and its epimers.  相似文献   

15.
Ketamine is widely used in medicine in combination with several benzodiazepines, including midazolam. The objectives of this study were to develop a novel HPLC‐MS/selected reaction monitoring (SRM) method capable of quantifying ketamine and norketamine using an isotopic dilution strategy in biological matrices and study the formation of norketamine, the principal metabolite of ketamine with and without the presence of midazolam, a well‐known CYP3A substrate. The chromatographic separation was achieved using a Thermo Betasil Phenyl 100 × 2 mm column combined with an isocratic mobile phase composed of acetonitrile, methanol, water and formic acid (60:20:20:0.4) at a flow rate of 300 μL/min. The mass spectrometer was operating in selected reaction monitoring mode and the analytical range was set at 0.05–50 μm . The precision (CV) and accuracy (NOM) observed were 3.9–7.8 and 95.9–111.1% respectively. The initial rate of formation of norketamine was determined using various ketamine concentrations and Km values of 18.4, 13.8 and 30.8 μm for rat, dog and human liver S9 fractions were observed, respectively. The metabolic stability of ketamine on liver S9 fractions was significantly higher in human (T1/2 = 159.4 min) compared with rat (T1/2 = 12.6 min) and dog (T1/2 = 7.3 min) liver S9 fractions. Moreover significantly lower IC50 and Ki values observed in human compared with rat and dog liver S9 fractions. Experiments with cDNA expressed CYP3A enzymes showed that the formation of norketamine is mediated by CYP3A but results suggest an important contribution from other isoenzymes, most likely CYP2C particularly in rat. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

16.
The folate analogue, 9‐thia‐5,10‐dideazafolic acid ( 3b ), was obtained in an efficient two‐step procedure in an overall yield of 60%. The previously unknown intermediate dimethyl‐thiocarbamic acid S‐(2‐amino‐3,4‐dihydo‐4‐oxo‐pyrido[2,3‐d]pyrimidin‐6‐yl) ester ( 5 ) was prepared via the condensation of 2,6‐diamino‐3H‐pyrimidin‐4‐one and S‐(2‐malonaldehyde)‐1,1,3,3‐tetramethylthiouronium bromide ( 4 ). Compound 5 , in a one pot procedure, was deprotected using sodium hydroxide and then coupled to diethyl N‐[(4‐chloromethyl)benzoyl]‐L‐glutamate, followed by saponification of the ethyl esters to give the 9‐thia‐5,10‐dideazafolic acid ( 3b ). Compound 3b was a potent inhibitor of human 5‐aminoimidazole‐4‐carboxamide ribonucleotide transformylase (Ki of 8 ± 5 μM) and showed no inhibition of human glycinamide ribonu‐cleotide transformylase at concentrations as high as 50 μM. Compound 3b was screened by the National Cancer Institute Developmental Therapeutics Program against 60 human tumors and was found to be active against a leukemia RPMI‐8226 cell line where the LC50 was 1 μM.  相似文献   

17.
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   

18.
The synthesis and evaluation as 5‐HT1A and 5‐HT7 serotonin receptor ligands of the two sets of O‐substituted hydroxybenzamides, structurally related to 2‐{3‐[4‐(2‐methoxyphenyl)piperazin‐1‐yl]propoxy}benzamide ( 1 ), (Ki 5‐HT1A = 21 nM, 5‐HT7 = 234 nM) are reported. To affect the affinity for 5‐HT1A and 5‐HT7 receptors, an amide moiety ( 2 , 3 , 4 , 5 , 6 ) and a hydrocarbon chain length ( 7 , 8 , 9 , 10 ) were modified. The serotonergic activity of compounds 2 , 3 , 4 , 5 , 6 , 7 , 8 , 9 , 10 was generally higher in the case of 5‐HT1A receptors compared with 5‐HT7 ones; the most active 5‐HT1A ligands being meta‐isomer 2 (Ki = 7 nM) and both analogs of 1 with the longest spacer, i.e., penta‐ and hexa‐methylene derivatives 9 and 10 (Ki = 4 and 3 nM, respectively). The observed biological properties of compounds 1 , 2 , 3 , 4 , 5 , 6 , 7 , 8 , 9 , 10 were elucidated using molecular modeling procedures. J. Heterocyclic Chem., (2010).  相似文献   

19.
4-deoxy-4-Phosphonomethyl-d-erythronate, an isosteric and hydrolytically stable analogue of the known ribose-5-phosphate isomerase inhibitor 4-deoxy-4-phospho-d-erythronate, was obtained by a 14-step synthesis from d-arabinose through an highly improved synthesis of the precursor 5-deoxy-5-phosphonomethyl-d-arabinose. The title compound appears as the first stable and potent competitive inhibitor of the enzyme catalyzed isomerization of ribose-5-phosphate to d-ribulose-5-phosphate (Ki=74 μM, Km/Ki=100), exhibiting only a 3-fold weaker inhibitory activity than its phosphate analogue.  相似文献   

20.
New drugs aimed at novel targets are urgently needed to combat the increasing rate of drug-resistant tuberculosis (TB). Herein, the National Cancer Institute Developmental Therapeutic Program (NCI-DTP) chemical library was screened against a promising new target, ketol–acid reductoisomerase (KARI), the second enzyme in the branched-chain amino acid (BCAA) biosynthesis pathway. From this library, 6-hydroxy-2-methylthiazolo[4,5-d]pyrimidine-5,7(4H,6H)-dione (NSC116565) was identified as a potent time-dependent inhibitor of Mycobacterium tuberculosis (Mt) KARI with a Ki of 95.4 nm . Isothermal titration calorimetry studies showed that this inhibitor bound to MtKARI in the presence and absence of the cofactor, nicotinamide adenine dinucleotide phosphate (NADPH), which was confirmed by crystal structures of the compound in complex with closely related Staphylococcus aureus KARI. It is also shown that NSC116565 inhibits the growth of H37Ra and H37Rv strains of Mt with MIC50 values of 2.93 and 6.06 μm , respectively. These results further validate KARI as a TB drug target and show that NSC116565 is a promising lead for anti-TB drug development.  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