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.
A new route via intermediate pseudoenantiomers was developed to synthesize racemic and enantiomerically pure new non‐peptidic inhibitors of thrombin, a key serine protease in the blood‐coagulation cascade. These ligands feature a conformationally rigid tricyclic core and are decorated with substituents to fill the major binding pockets (distal (D), proximal (P), selectivity (S1), and oxyanion hole) at the thrombin active site (Fig. 1). The key step in the preparation of the new inhibitors is the 1,3‐dipolar cycloaddition between an optically active azomethine ylide, prepared in situ from L ‐(4R)‐hydroxyproline and 4‐bromobenzaldehyde, and N‐piperonylmaleimide (Scheme 1). According to this protocol, tricyclic imide (compounds (±)‐ 15 ‐(±)‐ 18 and (+)‐ 21 ) and lactam (compound (+)‐ 2 ) inhibitors with OH or ether substituents at C(7) in the proline‐derived pyrrolidine ring were synthesized to specifically explore the binding features of the oxyanion hole (Schemes 2–4). Biological assays (Table) showed that the polar oxyanion hole in thrombin is not suitable for the accommodation of bulky substituents of low polarity, thereby confirming previous findings. In contrast, tricyclic lactam (+)‐ 2 (Ki=9 nM , Ki(trypsin)/Ki(thrombin)=1055) and tricyclic imide (+)‐ 21 (Ki=36 nM , Ki(trypsin)/Ki(thrombin)=50) with OH‐substituents at the (R)‐configured C(7)‐atom are among the most‐potent and most‐selective thrombin inhibitors in their respective classes, prepared today. While initial modeling predicted H‐bonding between the OH group at C(7) in (+)‐ 2 and (+)‐ 21 with the H2O molecule bound in the oxyanion hole (Fig. 2), the X‐ray crystal structure of the complex of (+)‐ 21 (Fig. 7, b) revealed a different interaction for this group. The propionate side chain of Glu192 undergoes a conformational change, thereby re‐orienting towards the OH group at C(7) under formation of a very short ionic H‐bond (O? H????OOC; d(O???O)=2.4 Å). The energetic contribution of this H‐bond, however, is negligible, due to its location on the surface of the protein and the unfavorable conformation of the H‐bonded propionate side chain.  相似文献   

2.
A novel class of nonpeptidic, active, and selective thrombin inhibitors has resulted from X‐ray‐structure‐based design and subsequent improvement of the initial lead molecules. These inhibitors possess a bi‐ or tricyclic central core structure with attached side chains to reach the three binding pockets (selectivity S1 pocket, distal D pocket, and proximal P pocket) present in the active site of the enzyme. The key step in the preparation of these compounds is the 1,3‐dipolar cycloaddition between an azomethine ylide, prepared in situ by the decarboxylative method from an aromatic aldehyde and an α‐amino acid, with an N‐substituted maleimide (e.g., see Schemes 1 and 2). All potent inhibitors contain an amidinium residue in the side chain for incorporation into the S1 pocket, which was introduced in the last step of the synthesis by a Pinner reaction. The compounds were tested in biological assays for activity against thrombin and the related serine protease trypsin. The first‐generation lead compounds (±)‐ 11 and (±)‐ 19 (Scheme 1) with a bicyclic central scaffold showed Ki values for thrombin inhibition of 18 μM and 0.67 μM , respectively. Conformationally more restricted second‐generation analogs (Scheme 2) were more active ((±)‐ 22i : Ki=90 nM (Table 1)); yet the selectivity for thrombin over trypsin remained weak. In the third‐generation compounds, a small lipophilic side chain for incorporation into the hydrophobic P pocket was introduced (Schemes 7 and 8). Since this pocket is present in thrombin but not in trypsin, an increase in binding affinity was accompanied by an increase in selectivity for thrombin over trypsin. The most selective inhibitor (Ki=13 nM , 760‐fold selectivity for thrombin over trypsin; Table 2) was (±)‐ 1 with an i‐Pr group for incorporation into the P pocket. Optical resolution of (±)‐ 1 (Scheme 9) provided (+)‐ 1 with a Ki value of 7 nM and a 740‐fold selectivity, whereas (−)‐ 1 was 800‐fold less active (Ki=5.6 μM , 21‐fold selectivity). The absolute configuration of the stronger‐binding enantiomer was assigned based on the X‐ray crystal structure of the complex formed between thrombin and this inhibitor. Compound (+)‐ 1 mimics the natural thrombin substrate, fibrinogen, which binds to the enzyme with the Ph group of a phenylalanine (piperonyl in (+)‐ 1 ) in the distal D pocket, with the i‐Pr group of a valine (i‐Pr in (+)‐ 1 ) in the proximal P pocket, and with a guanidinium side chain of an arginine residue (phenylamidinium group in (+)‐ 1 ) in the selectivity S1 pocket of thrombin. A series of analogs of (±)‐ 1 with the phenylamidinium group replaced by aromatic and aliphatic rings bearing OH or NH2 groups (Schemes 10 – 14) were not effectively bound by thrombin. A number of X‐ray crystal‐structure analyses of free inhibitors confirmed the high degree of preorganization of these compounds in the unbound state. Since all inhibitors prefer similar modes of association with thrombin, detailed information on the strength of individual intermolecular bonding interactions and their incremental contribution to the overall free energy of complexation was generated in correlative binding and X‐ray studies. The present study demonstrates that defined mutations in highly preorganized inhibitors provide an attractive alternative to site‐directed mutagenesis in exploring molecular‐recognition phenomena at enzyme active sites.  相似文献   

3.
A new class of nonpeptidic inhibitors of the ZnII‐dependent metalloprotease neprilysin with IC50 values in the nanomolar activity range (0.034–0.30 μM ) were developed based on structure‐based de novo design (Figs. 1 and 2). The inhibitors feature benzimidazole and imidazo[4,5‐c]pyridine moieties as central scaffolds to undergo H‐bonding to Asn542 and Arg717 and to engage in favorable ππ stacking interactions with the imidazole ring of His711. The platform is decorated with a thiol vector to coordinate to the ZnII ion and an aryl residue to occupy the hydrophobic S1′ pocket, but lack a substituent for binding in the S2′ pocket, which remains closed by the side chains of Phe106 and Arg110 when not occupied. The enantioselective syntheses of the active compounds (+)‐ 1 , (+)‐ 2 , (+)‐ 25 , and (+)‐ 26 were accomplished using Evans auxiliaries (Schemes 2, 4, and 5). The inhibitors (+)‐ 2 and (+)‐ 26 with an imidazo[4,5‐c]pyridine core are ca. 8 times more active than those with a benzimidazole core ((+)‐ 1 and (+)‐ 25 ) (Table 1). The predicted binding mode was established by X‐ray analysis of the complex of neprilysin with (+)‐ 2 at 2.25‐Å resolution (Fig. 4 and Table 2). The ligand coordinates with its sulfanyl residue to the ZnII ion, and the benzyl residue occupies the S1′ pocket. The 1H‐imidazole moiety of the central scaffold forms the required H‐bonds to the side chains of Asn542 and Arg717. The heterobicyclic platform additionally undergoes π‐π stacking with the side chain of His711 as well as edge‐to‐face‐type interactions with the side chain of Trp693. According to the X‐ray analysis, the substantial advantage in biological activity of the imidazo‐pyridine inhibitors over the benzimidazole ligands arises from favorable interactions of the pyridine N‐atom in the former with the side chain of Arg102. Unexpectedly, replacement of the phenyl group pointing into the deep S1′ pocket by a biphenyl group does not enhance the binding affinity for this class of inhibitors.  相似文献   

4.
5.
A new class of nonpeptidic inhibitors of the malarial aspartic protease plasmepsin II (PMII) with up to single‐digit micromolar activities (IC50 values) was developed by structure‐based de novo design. The active‐site matrix used in the design was based on an X‐ray crystal structure of PMII, onto which the major conformational changes seen in the structure of renin upon complexation of 4‐arylpiperidines – including the unlocking of a new hydrophobic (flap) pocket – were modeled. The sequence identity of 35% between mature renin and PMII had prompted us to hypothesize that an induced‐fit adaptation around the active site as observed in renin might also be effective in PMII. The new inhibitors contain a central 11‐azatricyclo[6.2.1.02,7]undeca‐2(7),3,5‐triene core, which, in protonated form, undergoes ionic H‐bonding with the two catalytic Asp residues at the active site of PMII (Figs. 1 and 2). This tricyclic scaffold is readily prepared by a Diels? Alder reaction between an activated pyrrole and a benzyne species generated in situ (Scheme 1). Two substituents with naphthyl or 1,3‐benzothiazole moieties are attached to the central core (Schemes 14) for accommodation in the hydrophobic flap and S1/S3 (or S2′, depending on the optical antipode of the inhibitor) pockets at the active site of the enzyme. The most‐potent inhibitors (±)‐ 19a – 19c (IC50 3–5 μM ) and (±)‐ 23b (2 μM ) (Table) bear an additional Cl‐atom on the 1,3‐benzothiazole moiety to fully fill the rear of the flap pocket. Optimization of the linker between the tricyclic scaffold and the 1,3‐benzothiazole moiety, based on detailed conformational analysis (Figs. 3 and 4), led to a further small increase in inhibitory strength. The new compounds were also tested against other aspartic proteases. They were found to be quite selective against renin, while the selectivity against cathepsin D and E, two other human aspartic proteases, is rather poor (Table). The detailed SARs established in this investigation provide a valuable basis for the design of the next generations of more‐potent and ‐selective PMII inhibitors with potential application in a new antimalarial therapy.  相似文献   

6.
(−)‐ 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   

7.
8.
Treatment of the unsaturated bicyclic homoallylic alcohols (E)‐ and (Z)‐ 5 and (Z)‐ and (E)‐ 10 and allenic alcohol 16 with an excess of ClSO3H in 2‐nitropropane or CH2Cl2 at − 80° afforded, in moderate yields (ca. 30–70%), diastereoisomer mixtures of racemic tetracyclic ethers 12a – c (Table 1) and 17a,b (Table 2), respectively. These kinetically controlled stereospecific transformations are believed to proceed via concerted or nonconcerted pathways (see Schemes 4 and 6) and the results are fully consistent with our earlier work. Representing novel didehydro bridged analogues of known, olfactively active labdane tricyclic ethers, the organoleptic properties of 12a – c and 17a,b are briefly described, especially those of 12c which, in the context of structure–activity studies, is a racemic didehydro analogue of the known ambergris odorant Ambrox®.  相似文献   

9.
Substituted phenyl‐N‐butyl carbamates ( 1 ) and p‐nitrophenyl‐N‐substituted carbamates ( 2 ) are characterized as “pseudo‐pseudo‐substrate” inhibitors of acetylcholinesterase. Since the inhibitors protonate in pH 7.0 buffer solution, the virtual inhibition constants (Ki's) of the protonated inhibitors can be calculated from the equation, ‐logKi' = ‐logKi ‐ pKa + 14. The ‐logKi' and logkc values for acetylcholinesterase inhibitions by carbamates 1 correlate with the Hammett equation (log(k/k0) = ρσ); moreover, those by carbamates 2 correlate with the Taft equation (log(k/k0) = ρ* σ*). With modified Hammett‐Taft cross‐interaction variations, multiple linear regressions of the ‐logKi' and logkc values of carbamates 1 and 2 give good correlations, and the cross‐interaction constants (ρXR) are 0.5 and 0.0, respectively. The ρXR value of 0.5 indicates that the carbamate O‐C(O)‐N‐R geometries for the transition states that lead to enzyme‐carbamate tetrahedral intermediates are all pseudo‐trans conformations. Therefore, the carbamate moiety of the inhibitors stretches along the active site gorge of the enzyme but does not bind in the acyl binding site pocket of the enzyme. Overall, the carbamate O‐C(O)‐N‐R geometries for carbamates 1 and 2 , protonated carbamates 1 and 2 , and the tetrahedral intermediate are all retained in pseudo‐trans conformations. The ρXR value of 0.0 suggests that the transition states that lead to the carbamyl enzymes are breaking C‐O bonds and are excluding the leaving groups, substituted phenols.  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.
Vilsmeier–Haack‐type cyclization of 1H‐indole‐4‐propanoic acid derivatives was examined as model construction for the A–B–C ring system of lysergic acid ( 1 ). Smooth cyclization from the 4 position of 1H‐indole to the 3 position was achieved by Vilsmeier–Haack reaction in the presence of K2CO3 in MeCN, and the best substrate was found to be the N,N‐dimethylcarboxamide 9 (Table 1). The modified method can be successfully applied to an α‐amino acid derivative protected with an N‐acetyl function, i.e., to 27 (Table 2); however, loss of optical purity was observed in the cyclization when a chiral substrate (S)‐ 27 was used (Scheme 5). On the other hand, the intramolecular Pummerer reaction of the corresponding sulfoxide 20 afforded an S‐containing tricyclic system 22 , which was formed by a cyclization to the 5 position (Scheme 3).  相似文献   

12.
The isoquinuclidines 7 and 8 were synthesised and tested as inhibitors of hexosaminidases from jack beans and from bovine kidney. These isoquinuclidines mimick the 1,4B‐conformer of a N‐acetyl‐glucosamine‐derived β‐d‐ glucopyranoside; they are competitive inhibitors with Ki values from 0.014 to 0.30 μM . The strong inhibition of these enzymes agrees with the hypothesis that the enzymatic hydrolysis of 2‐acetamido‐2‐deoxy‐β‐d‐ glucopyranosides proceeds via a boat‐like conformer with a pseudo‐axial scissile glycosidic bond and a pseudo‐axial acetamido substituent optimally oriented to effect an intramolecular substitution of the aglycon.  相似文献   

13.
The Pseudomonas species lipase inhibition shows enantioselectivity for R‐enantiomer over S‐enantiomer of exo‐2‐norbornyl‐Nn‐butylcarbamates. R‐, S‐, and racemic‐exo‐2‐norbornyl‐Nn‐butylcarbamates are all characterized as pseudo substrate inhibitors of the enzyme. Thus, the mechanism for Pseudomonas species lipase‐catalyzed hydrolysis of the inhibitor is formation of the first enzyme‐inhibitor Michaelis complex via nucleophilic attack of the active site serine to the inhibitor (Ki step) then formation of the butylcarbamyl enzyme intermediate from this complex (k2 step). Comparison of bimolecular rate constants (ki = k2 / Ki) of the inhibitors indicates that R‐enantiomer is 1.8 times more potent than S‐enantiomer. Thus, Pseudomonas species lipase shows enantioselectivity of 1.8 for Rexo‐2‐norbornyl‐Nn‐butyl‐carbamate over Sexo‐2‐norbornyl‐Nn‐butylcarbamate. Protein‐ligand interaction studies on both enantiomers of exo‐2‐norbornyl‐Nn‐butylcarbamate as inhibitors of Pseudomonas species lipase using AutoDock suggest that R‐enantiomer binds more tightly into the active site of the enzyme than S‐enantiomer. The norbornyl ring of Sexo‐2‐norbornyl‐Nn‐butylcarbamate is repulsive to Ser 82 and His 251 of the catalytic triad as well as to Met 16 of the oxyanion hole. These repulsions may create few unfavorable interactions between Sexo‐2‐norbornyl‐Nn‐butylcarbamate and the enzyme and make this inhibitor a less potent one.  相似文献   

14.
Inhibition of the enzyme catechol O‐methyltransferase (COMT) represents a viable strategy for regulation of the catabolism of catecholamine neurotransmitters or their precursors, and is of considerable interest in the therapy of Parkinson's disease. Herein, we report the development of a new generation of potent bisubstrate inhibitors of COMT derived from nitro‐substituted ligand 1 (Ki = 28 nM , Table 1), which achieve high biological activity despite the lack of a NO2 substituent on the catechol moiety. Their synthesis takes advantage of a convergent approach, in which a series of functionalized catechol intermediates is prepared (Schemes 2–7) and coupled to a common adenosine‐derived allylic amine building block (Scheme 8). Biological activities of the newly synthesized inhibitors, determined by in vitro enzymatic assay and kinetic studies, clearly demonstrate that high inhibitory potency of the bisubstrate inhibitors is not correlated with the pKa of the catechol OH groups. Aromatic residues, connected to the catechol via a biaryl‐type linkage, were found to maximally benefit from additional favorable hydrophobic interactions with the enzyme and thus to be preferred replacements of the NO2 group in 1 . A competitive kinetic inhibition mechanism (Fig. 2) with respect to the cofactor binding site was confirmed in all cases, supporting a bisubstrate inhibition mode for inhibitors 2 – 19 .  相似文献   

15.
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   

16.
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   

17.
For substituted phenyl‐N‐butyl carbamates (1) and 4‐nitrophenyl‐N‐substituted carbamates (2), linear relationships between values of NH proton chemical shift (δNH), pKa, and logk[OH] and Hammett substituent constant (σ) or Taft substituent constant (σ*) are observed. Carbamates 1 and 2 are pseudo‐substrate inhibitors of porcine pancreatic cholesterol esterase. Thus, the mechanism of the reaction necessitates that the inhibitor molecule and the enzyme form the enzyme‐inhibitor tetrahedral species at the Ki step of the reaction and then form the carbamyl enzyme at the kc step of the reaction. Linear relationships between the logarithms of Ki and kc for cholesterol esterase by carbamates 1 and σ are observed, and the reaction constants (ρs) are ?3.4 and ?0.13, respectively. Therefore, the above reaction forms the negative‐charge tetrahedral species and follows the formation of the relatively neutral carbamyl enzymes. For the inhibition of cholesterol esterase by carbamates 2 except 4‐nitrophenyl‐N‐phenyl carbamate and 4‐nitrophenyl‐N‐t‐butyl carbamate, linear relationships of ‐logKi and logkc with σ* are observed and the ρ* values are ?0.50 and 1.03, respectively. Since the above reaction also forms the negative‐charge tetrahedral intermediate, it is possible that the Ki step of this reaction is further divided into two steps. The first Ki step is the development of the positive‐charge at the carbamate nitrogen from the protonation of the carbamate nitrogen. The second Ki step is the formation of the tetrahedral intermediate with the negative‐charge at the carbonyl oxygen. From Arrhenius plots of a series of inhibition reactions by carbamates 1 and 2, the isokinetic and isoequilibrium temperatures are different from the reaction temperature (25°C). Therefore, the observed ρ and ρ* values only depend upon the electronic effects of the substituents. Taken together, the cholesterol esterase inhibition mechanism by carbamates 1 and 2 is proposed.  相似文献   

18.
Methods have been developed for the synthesis of 7α‐amino‐ and 7α‐(aminomethyl)‐N‐cyclopropylmethyl‐6,14‐endo‐ethanotetrahydronororipavines and their cinnamoyl derivatives (Schemes 1 and 3). In displacement binding assays, the cinnamoyl derivatives 4c and 5c had high affinity for opioid receptors, but no particular selectivity for any receptor type or differences in affinity between 4c and 5c (Table 1). In tissue assays for opioid receptor function, in which both 4c and 5c were potent antagonists, the aminomethyl derivative 5c was 20‐ to 70‐fold more potent than the amino derivative 4c (Table 2). These data are in keeping with previously reported in vivo data and confirm the major effect of the methylene spacer in 5c .  相似文献   

19.
Plasmepsin II (PMII), a malarial aspartic protease involved in the catabolism of hemoglobin in parasites of the genus Plasmodium, and renin, a human aspartic protease, share 35% sequence identity in their mature chains. Structures of 4‐arylpiperidine inhibitors complexed to human renin were reported by Roche recently. The major conformational changes, compared to a structure of renin, with a peptidomimetic inhibitor were identified and subsequently modeled in a structure of PMII (Fig. 1). This distorted structure of PMII served as active‐site model for a novel class of PMII inhibitors, according to a structure‐based de novo design approach (Fig. 2). These newly designed inhibitors feature a rigid 7‐azabicyclo[2.2.1]heptane scaffold, which, in its protonated form, is assumed to undergo ionic H‐bonding with the two catalytic Asp residues at the active site of PMII. Two substituents depart from the scaffold for occupancy of either the S1/S3 or S2′‐pocket and the hydrophobic flap pocket, newly created by the conformational changes in PMII. The inhibitors synthesized starting from N‐Boc‐protected 7‐azabicyclo[2.2.1]hept‐2‐ene ( 6 ; Schemes 15) displayed up to single‐digit micromolar activity (IC50 values) toward PMII and good selectivity towards renin. The clear structure? activity relationship (SAR; Table) provides strong validation of the proposed conformational changes in PMII and the occupancy of the resulting hydrophobic flap pocket by our new inhibitors.  相似文献   

20.
Treatment of the acyclic tetraenols (E)‐ and (Z)‐ 2 with an excess of ClSO3H in 2‐nitropropane at ? 80° stereoselectively afforded in 30 and 43% yield, respectively, diastereoisomer mixtures of the racemic, tricyclic ethers 1c,d and 1a,b , together with 20 (Table). Under identical conditions, but with the acyclic pentaenol 10 (1 : 1 diastereoisomer mixture) as substrate, the tricyclic ethers 22a / 22b (10 : 1) were isolated in 27% yield. These kinetically controlled stereospecific transformations are thought to proceed via non‐concerted pathways (see Schemes 5 and 7), fully consistent with our earlier work. In contrast, another set of reaction conditions (CF3CO2H, CH2Cl2, ? 15° to ? 10°) was used for the cyclization of the monocyclic dienols (E)‐ 3 and (Z)‐ 3 , which resulted in the non‐stereoselective formation of the major products 1c,d and 1a,b , respectively, in 35–37% yield. Representing novel didehydro analogues of the known ambergris odorant (±)‐Ambrox® and its diastereoisomers, the qualitative organoleptic properties of 1a – d and of the 10 : 1 diastereoisomer mixture of the novel tetradehydro analogues 22a / 22b are briefly described.  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