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.
采用多拷贝同时搜寻方法(MCSS)分析得到了CaNMT活性位点的疏水区域、氢键结合位点和负电性区域. MCSS计算结果显示, CaNMT活性位点有两个疏水性比较强的区域: 一个由Tyr107, Tyr109, Val108, Phe117, Phe123, Ala127, Phe176和Leu337等残基组成; 另一个由Phe115, Phe240和Phe339组成. CaNMT活性位点发现有两个氢键作用区域, 其中Tyr119, His227, Asn392和Leu451是与已有抑制剂的氢键结合位点, Tyr107, Asn175, Thr211和Asp412是新发现的氢键结合位点, 而且在NMT家族中高度稳定, 它们对设计新结构类型的CaNMT抑制剂具有重要作用. Leu451是负电性兼氢键作用位点, 是抑制剂设计时所必需考虑的位点.  相似文献   

2.
VEGFR2介导肿瘤诱导的血管生成作用, 是抑制肿瘤生长和转移的新靶点. 为深入探讨VEGFR2活性腔性质以及与抑制剂的结合模式, 采用多拷贝同时搜寻法(MCSS)研究VEGFR2活性腔的性质, 然后用分子对接方法对5个已上临床的VEGFR抑制剂与VEGFR2活性腔进行对接计算, 讨论它们的结合模式, 确定与配体结合相关的关键残基. 研究发现: 疏水腔I, II是配体结合的关键区域, 残基Glu915, Cys917是关键的氢键作用位点, Lys866, Glu883和Asp1044形成的极性区域对提高配体亲合力很重要, 疏水腔III和极性腔IV是额外增强配体结合力的区域, IV区的Arg1030可提供额外的氢键作用位点. 本研究可为全新VEGFR2抑制剂的合理药物设计提供理论依据, 为寻找新的抗肿瘤药物奠定基础.  相似文献   

3.
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   

4.
Docking calculations that allow the estimation of the binding energy of small ligands in the GIIA sPLA2 active site were used in a structure-based design protocol. Four GIIA sPLA2 inhibitors co-crystallised with the enzyme, were used for examining the enzyme active site and for testing the FlexX in SYBYL 6.8 molecular docking program to reproduce the crystallographic experimental data. The FPL67047XX inhibitor was chosen as a prototype structure for applying free energy perturbation (FEP) studies. Structural modifications of the initial structure of the FPL67047XX inhibitor (IC50 0.013 μM) were performed in an effort to optimise the interactions in the GIIA sPLA2 active site. The structural modifications were based on: (1) the exploration of absolute configuration (i.e. comparison of the binding score of (R)- and (S)-enantiomers); (2) bioisosterism (i.e. replacement of the carboxylate group with the bioisosteric sulphonate and phosphonate groups); (3) insertion of substituents that fit better in the active site. The generated new structures exhibited higher binding energy. Such structures may spark off the interest of medicinal chemists for synthesizing potentially more active GIIA sPLA2 inhibitors.  相似文献   

5.
A homology model of Mycobacterium avium complex dihydrofolate reductase (MAC DHFR) was constructed on the basis of the X-ray crystal structure of Mycobacterium tuberculosis (Mtb) DHFR. The homology searching of the MAC DHFR resulted in the identification of the Mtb DHFR structure (PDB 1DF7) as the template for the model building. The MAC enzyme sequence was aligned to that of the Mtb counterpart using a modified Needleman and Wunsch methodology. The initial geometry to be modeled was copied from the template, either fully or partially depending on whether the residues were conserved or not, respectively. Using a randomized modeling procedure, 10 independent models of the target protein were built. The cartesian average of all the model structures was then refined using molecular mechanics. The resulting model was assessed for stereochemical quality using a Ramachandran plot and by analyzing the consistency of the model with the experimental data. The structurally and functionally important residues were identified from the model. Further, 5-deazapteridines recently reported as inhibitors of MAC DHFR were docked into the active site of the developed model. All the seven inhibitors used in the docking study have a similar docking mode at the active site. The network of hydrogen bonds around the 2,4-diamino-5-deazapteridine ring was found to be crucial for the binding of the inhibitors with the active site residues. The 5-methyl group of the inhibitors was located in a narrow hydrophobic pocket at the bottom of the active site. The relative values of the three torsion angles of the inhibitors were found to be important for the proper orientation of the inhibitor functional groups into the active site.  相似文献   

6.
The design of novel organic compounds containing no strongly basic amidine or guanidine functional groups typical of serine protease inhibitors was performed to develop an oral anticoagulant drug. A three-dimensional computational model for thrombin active site was constructed and optimized for docking of small-molecule organic compounds and calculating the energies of inhibitor-enzyme interactions. Novel racemic derivatives of 1-[2-(4-chlorophenylthio)acetyl]-5-phenylpyrrolidine-2,4-dicarboxylic acids were synthesized for which Cl-π interactions between the inhibitors and the S1 pocket of thrombin active site are predicted by modeling. The compounds synthesized deactivate thrombin in vitro and the inhibition properties show good correlations with the results of calculations.  相似文献   

7.
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   

8.
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   

9.
Theileria annulata secretes peptidyl prolyl isomerase enzyme (TaPIN1) to manipulate the host cell oncogenic signaling pathway by disrupting the tumor suppressor F-box and WD repeat domain-containing 7 (FBW7) protein level leading to an increased level of c-Jun proto-oncogene. Buparvaquone is a hydroxynaphthoquinone anti-theilerial drug and has been used to treat theileriosis. However, TaPIN1 contains the A53 P mutation that causes drug resistance. In this study, potential TaPIN1 inhibitors were investigated using a library of naphthoquinone derivatives. Comparative models of mutant (m) and wild type (wt) TaPIN1 were predicted and energy minimization was followed by structure validation. A naphthoquinone (hydroxynaphthalene-1,2-dione, hydroxynaphthalene-1,4-dione) and hydroxynaphthalene-2,3-dione library was screened by Schrödinger Glide HTVS, SP and XP docking methodologies and the docked compounds were ranked by the Glide XP scoring function. The two highest ranked docked compounds Compound 1 (4-hydroxy-3-[3,4,5-trihydroxy-6-(hydroxymethyl)oxan-2-yl]oxynaphthalene-1,2-dione) and Compound 2 (6-acetyl-1,4,5,7,8-pentahydroxynaphthalene-2,3-dione) were used for further molecular dynamics (MD) simulation studies. The MD results showed that ligand Compound 1 was located in the active site of both mTaPIN1 and wtTaPIN1 and could be proposed as a potential inhibitor by acting as a substrate antagonist. However, ligand Compound 2 was displaced away from the binding pocket of wtTaPIN1 but was located near the active site binding pocket of mTaPIN1 suggesting that could be selectively evaluated as a potential inhibitor against the mTaPIN1. Compound 1 and Compound 2 ligands are potential inhibitors but Compound 2 is suggested as a better inhibitor for mTaPIN1. These ligands could also further evaluated as potential inhibitors against human peptidyl prolyl isomerase which causes cancer in humans by using the same mechanism as TaPIN1.  相似文献   

10.
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   

11.
Free energy perturbation calculations were performed to determine the free energy of binding associated with the presence of perhaps an unusual hydroxyl group in the transition state analog of nebularine, an inhibitor of the enzyme adenosine deaminase. The presence of a single hydroxyl group in this inhibitor has been found to contribute ?9.8 kcal/mol to the free energy of binding, with a 108-fold increase in the binding affinity by the enzyme. In this work, we calculate the difference in solvation free energy for the 1,6-dihydropurine complex versus that of the 6-hydroxyl-1,6-dihydropurine complex to determine if this marked increase in binding affinity is attributed to an unusually hydrophobic hydroxyl group. The calculated ΔG associated for the solvation free energy is ?11.8 kcal/mol. This large change in the solvation free energy suggests that this hydroxyl is instead unusually hydrophilic and that the difference in free energy of interaction for the two inhibitors to the enzyme must be at least ca. 20 kcal/mol. Although the crystal structure for adenosine deaminase is currently not known, we attempt to mimic the nature of the active site by constructing models which simulate the enzyme-inhibitor complex. We present a first attempt at determining the change in free energy of binding for a system in which structural data for the enzyme is incomplete. To do this, we construct what we believe is a minimal model of the binding between adenosine deaminase and an inhibitor. The active site is simulated as a single charged carboxyl group which can form a hydrogen bond with the hydroxyl group of the analog. Two different carboxyl anion models are used. In the first model, the association is modeled between an acetic acid anion and the modified inhibitor. The second model consists of a hydrophobic amino acid pocket with an interior Glu residue in the active site. From these models we calculate the change in free energy of association and the overall change in free energy of binding. We calculate the free energies of interaction both in the absence and presence of water. We conclude from this that the presence of a single suitably placed-CO?2 group probably cannot explain the binding effect of the-OH group and that additional interactions will be found in the adenosine deaminase active site.  相似文献   

12.
Functionality map analysis of the active site cleft of human thrombin   总被引:1,自引:0,他引:1  
Summary The Multiple Copy Simultaneous Search methodology has been used to construct functionality maps for an extended region of human thrombin, including the active site. This method allows the determination of energetically favorable positions and orientations for functional groups defined by the user on the three-dimensional surface of a protein. The positions of 10 functional group sites are compared with those of corresponding groups of four thrombin-inhibitor complexes. Many, but not all features, of known thrombin inhibitors are reproduced by the method. The results indicate that certain aspects of the binding modes of these inhibitors are not optimal. In addition, suggestions are made for improving binding by interaction with functional group sites on the thrombin surface that are not used by the thrombin inhibitors. Abbreviations: MCSS, multiple copy simultaneous search; PPACK, d-phenylalanyl-l-propyl-l-arginine chloromethane; NAPAP, N -(2-naphthylsulfonylglycyl)-d-para-amidinophenylalanylpiperidine; argatroban, (2R,4R)-4-methyl-1-[N -(3-methyl-1,2,3,4-tetrahydro-8-quinolinylsulfonyl)-l-arginyl]-2-piperidine carboxylic acid; rms, root mean square. The thrombin residues are numbered according to the chymotrypsin-based numbering by Bode et al. [8]. P1, P2, P3, etc., denote the peptide inhibitor residues on the amino-terminal side of the scissile peptide bond, and S1, S2, S3, etc., the corresponding subsites of thrombin  相似文献   

13.
14.
The enzymes of the non‐mevalonate pathway for isoprenoid biosynthesis have been identified as attractive targets with novel modes of action for the development of herbicides for crop protection and agents against infectious diseases. This pathway is present in many pathogenic organisms and plants, but absent in mammals. By using high‐throughput screening, we identified highly halogenated marine natural products, the pseudilins, to be inhibitors of the third enzyme, IspD, in the pathway. Their activity against the IspD enzymes from Arabidopsis thaliana and Plasmodium vivax was determined in photometric and NMR‐based assays. Cocrystal structures revealed that pseudilins bind to an allosteric pocket by using both divalent metal ion coordination and halogen bonding. The allosteric mode of action for preventing cosubstrate (CTP) binding at the active site was elucidated. Pseudilins show herbicidal activity in plant assays and antiplasmodial activity in cell‐based assays.  相似文献   

15.

Abstract  

The similarity between Plasmodium falciparum phosphodiesterase enzymes (PfPDEs) and their human counterparts have been examined and human PDE9A was found to be a suitable template for the construction of homology models for each of the four PfPDE isoforms. In contrast, the architecture of the active sites of each model was most similar to human PDE1. Molecular docking was able to model cyclic guanosine monophosphate (cGMP) substrate binding in each case but a docking mode supporting cyclic adenosine monophosphate (cAMP) binding could not be found. Anticipating the potential of PfPDE inhibitors as anti-malarial drugs, a range of reported PDE inhibitors including zaprinast and sildenafil were docked into the model of PfPDEα. The results were consistent with their reported biological activities, and the potential of PDE1/9 inhibitor analogues was also supported by docking.  相似文献   

16.
Summary We have developed an approach to search for molecules that can be used as lead compounds in designing an inhibitor for a given proteolytic enzyme when the 3D structure of a homologous protein is known. This approach is based on taking the cast of the binding pocket of the protease and comparing its dimensions with that of the dimensions of small molecules. Herein the 3D structure of papain is used to model cathepsin L using the comparative modeling technique. The cast of the binding pocket is computed using the crystal structure of papain because the structures of papain and the model of cathepsin L are found to be similar at the binding site. The dimensions of the cast of the binding site of papain are used to screen for molecules from the Cambridge Structural Database (CSD) of small molecules. Twenty molecules out of the 80 000 small molecules in the CSD are found to have dimensions that are accommodated by the papain binding pocket. Visual comparison of the shapes of the cast and the 20 screened molecules resulted in identifying brevotoxin b, a toxin isolated from the red tide dinoflagellate Ptycho brevis (previously classified as Gymonodium breve), as the structure that best fits the binding pocket of papain. We tested the proteolytic activity of papain and cathepsin L in the presence of brevotoxin b and found inhibition of papain and cathepsin L with Kis of 25 M and 0.6 M, respectively. We also compare our method with a more elaborate method in the literature, by presenting our results on the computer search for inhibitors of the HIV-1 protease.  相似文献   

17.
Structure‐activity relationships for new members of a class of nonpeptidic, low‐molecular‐weight inhibitors of thrombin, a key serine protease in the blood coagulation cascade, are described. These compounds, which originate from X‐ray‐structure‐based design, feature a conformationally rigid, bi‐ or tricyclic core from which side chains diverge into the four major binding pockets (distal D, proximal P, recognition or specificity S1, and oxyanion hole O) at the thrombin active site (Fig. 1). Phenylamidinium is the side chain of choice for the S1 pocket, while the most active inhibitors orient an i‐Pr group into the P‐pocket (Table 1). The key step in the synthesis of the inhibitors is the construction of the central bi‐ or tricyclic scaffold by 1,3‐dipolar cycloaddition of an in situ prepared azomethine ylide and an N‐substituted maleimide (Schemes 1–3, and 8–10). One series of compounds was designed to explore the binding features of the large hydrophobic D pocket. This pocket provides space for lipophilic residues as bulky as benzhydryl groups. A new strategy was developed, allowing introduction of these sterically demanding substituents very late in the synthesis (Schemes 5 and 6). Benzhydryl derivative (±)‐ 2 was found to be the most selective member (Ki (trypsin)/Ki (thrombin)=1200) of this class of nonpeptidic thrombin inhibitors, while the ‘dipiperonyl' analog (±)‐ 3 (Ki=9 nM , 7.60‐fold selectivity) displays the highest potency of all compounds prepared so far (Table 1). A second series of inhibitors features side chains designed to orient into the oxyanion hole and to undergo H‐bonding with the backbone NH groups lining the catalytic site of the enzyme. Unfortunately, neither activity nor selectivity could be substantially improved by introduction of these substituents (Table 2). Presumably, the high degree of pre‐organization and the rigidity of the tightly bound scaffolds prevents the new substituents from assuming a position that would allow favorable interactions in the oxyanion hole. However, the oxyanion hole and the S1′ pocket next to it were found to be capable of accommodating quite large groups, which leaves much room for further exploration.  相似文献   

18.
《中国化学快报》2023,34(7):108072
New Delhi metallo-β-lactamase 1 (NDM-1) can hydrolyze most β-lactam antibiotics, which is the major factor for drug resistance of Gram-negative bacteria. The binding of most reversible inhibitors to NDM-1 is relatively weak due to the shallow active pocket of NDM-1. Alternatively, irreversible covalent inhibitors can prevent their dissociation from the target, leading to permanent inactivation of the protein. Herein, we report a series of irreversible covalent inhibitors of NDM-1 targeting the conserved Lys211 in the active pocket. Several methods, including mass spectrometry, sodium dodecyl sulfate-polyacrylamide gel electrophoresis, fluorescent labeling, and coumarin probe were used to demonstrate that pentafluorophenyl ester formed a covalent bond with Lys211. Moreover, our target inhibitor, in combination with meropenem, achieved an antibacterial effect on drug-resistant bacteria, along with an excellent safety profile. Our new strategy in designing lysine-targeted irreversible covalent NDM-1 inhibitors provides a potential option for the clinical treatment of Gram-negative bacteria.  相似文献   

19.
Cyclooxygenase (COX) and lipoxygenase (LOX) are key targets for the development of new anti-inflammatory agents. LOX, which is involved in the biosynthesis of mediators in inflammation and allergic reactions, was selected for a biochemical screening campaign to identify LOX inhibitors by employing the main natural product library of Brazilian biodiversity. Two prenyl chalcones were identified as potent inhibitors of LOX-1 in the screening. The most active compound, (E)-2-O-farnesyl chalcone, decreased the rate of oxygen consumption to an extent similar to that of the positive control, nordihydroguaiaretic acid. Additionally, studies on the mechanism of the action indicated that (E)-2-O-farnesyl chalcone is a competitive LOX-1 inhibitor. Molecular modeling studies indicated the importance of the prenyl moieties for the binding of the inhibitors to the LOX binding site, which is related to their pharmacological properties.  相似文献   

20.
One approach to combinatorial ligand design begins by determining optimal locations (i.e., local potential energy minima) for functional groups in the binding site of a target macromolecule. MCSS and GRID are two methods, based on significantly different algorithms, which are used for this purpose. A comparison of the two methods for the same functional groups is reported. Calculations were performed for nonpolar and polar functional groups in the internal hydrophobic pocket of the poliovirus capsid protein, and on the binding surface of the src SH3 domain. The two approaches are shown to agree qualitatively; i.e., the global characteristics of the functional group maps generated by MCSS and GRID are similar. However, there are significant differences in the relative interaction energies of the two sets of minima, a consequence of the different functional form used to evaluate polar interactions (electrostatics and hydrogen bonding) in the two methods. The single sphere representation used by GRID affords only positional information, supplemented by the identification of hydrogen bonding interactions. By contrast, the multi-atom representation of most MCSS groups yields in both positional and orientational information. The two methods are most similar for small functional groups, while for larger functional groups MCSS yields results consistent with GRID but superior in detail. These results are in accord with the somewhat different purposes for which the two methods were developed. GRID has been used mainly to introduce functionalities at specific positions in lead compounds, in which case the orientation is predetermined by the structure of the latter. The orientational information provided by MCSS is important for its use in the de novo design of large, multi-functional ligands, as well as for improving lead compounds.  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