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.
The stacking interaction between a tyrosine residue and the sugar ring at the catalytic subsite -1 is strictly conserved in the glycoside hydrolase family 13 enzymes. Replacing Tyr100 with leucine in cyclodextrin glycosyltransferase (CGTase) from Bacillus sp. 1011 to prevent stacking significantly decreased all CGTase activities. The adjacent stacking interaction with both Phe183 and Phe259 onto the sugar ring at subsite +2 is essentially conserved among CGTases. F183L/F259L mutant CGTase affects donor substrate binding and/or acceptor binding during transglycosylation [Nakamura et al. (1994) Biochemistry 33, 9929-9936]. To elucidate the precise role of carbohydrate/aromatic stacking interaction at subsites -1 and +2 on the substrate binding of CGTases, we analyzed the X-ray structures of wild-type (2.0 A resolution), and Y100L (2.2 A resolution) and F183L/F259L mutant (1.9 A resolution) CGTases complexed with the inhibitor, acarbose. The refined structures revealed that acarbose molecules bound to the Y100L mutant moved from the active center toward the side chain of Tyr195, and the hydrogen bonding and hydrophobic interaction between acarbose and subsites significantly diminished. The position of pseudo-tetrasaccharide binding in the F183L/F259L mutant was closer to the non-reducing end, and the torsion angles of glycosidic linkages at subsites -1 to +1 on molecule 1 and subsites -2 to -1 on molecule 2 significantly changed compared with that of each molecule of wild-type-acarbose complex to adopt the structural change of subsite +2. These structural and biochemical data suggest that substrate binding in the active site of CGTase is critically affected by the carbohydrate/aromatic stacking interaction with Tyr100 at the catalytic subsite -1 and that this effect is likely a result of cooperation between Tyr100 and Phe259 through stacking interaction with substrate at subsite +2.  相似文献   

2.
The crystal structure of the ternary complex of NADPH, the potent antifolate [2, 4-diamino-5-?3-[3-(2-acetyloxyethyl)-3-benzyltriazen-1-yl]-4 -chloroph enyl?-6-ethylpyrimidine] (TAB, 1) and Pneumocystis carinii dihydrofolate reductase (pcDHFR), refined to 2.1 A resolution, reveals that TAB binds similar to the antifolates trimethoprim and methotrexate. These data also reveal multiple conformations for the binding geometry of TAB with two preferred orientations of the acetyloxy and benzyl groups that results from a 180 degrees rotation about the N2-N3 triazenyl bond. The methyl of the acetyloxy and benzyl ring of TAB probes large hydrophobic regions of the p-aminobenzoyl folate binding pocket of the active site, in particular the region near Phe69, which is unique to the pcDHFR sequence. These results confirm prior molecular modeling investigations of the binding of TAB to pcDHFR that identified four low-energy binding geometries, two involving rotations about the terminal N(2)-N(3) triazenyl linkage and two involving atropisomerism about the pivotal pyrimethamine-phenyl bond. The primary differences in the molecular dynamics (MD) models and those observed in this crystal complex result from small conformational changes in active-site residues on energy minimization. However, two MD models place the acetyloxy and benzyl ring groups in a region of the active site between the cofactor-binding region and the p-aminobenzoyl folate pocket; an orientation never observed in any DHFR crystal structure to date. These conformers interact with solvent near the enzyme surface and are probably not observed due to the loss of specific hydrogen bonds with the enzyme. The high species pcDHFR selectivity of TAB could be the result of ligand flexibility that enables multiple binding orientations at the enzyme active site. Further modification of the acetyloxy region of TAB could increase its potency and selectivity for pcDHFR.  相似文献   

3.
Six 2,4-diaminopyrido[2,3-d]pyrimidines with a 6-methylthio bridge to an aryl group were synthesized and biologically evaluated as inhibitors of Pneumocystis carinii (pc) and Toxoplasma gondii (tg) dihydrofolate reductase (DHFR). The syntheses of analogues 3-8 were achieved by nucleophilic displacement of 2,4-diamino-6-bromomethylpyrido[2,3-d]pyrimidine 14 with various arylthiols. The alpha-naphthyl analogue 4 showed the highest selectivity ratios of 3.6 and 8.7 against pcDHFR and tgDHFR, respectively, versus rat liver (rl) DHFR. The beta-naphthyl analogue 5 exhibited the highest potency within the series with an IC(50) value against pcDHFR and tgDHFR of 0.17 and 0.09 microM, respectively. Analogue 4 was evaluated for in vitro antimycobacterium activity and was shown to inhibit the growth of Mycobacterium tuberculosis H(37)Rv cells by 58% at a concentration of 6.25 microg/mL.  相似文献   

4.
Molecular docking is routinely used for understanding drug‐receptor interaction in modern drug design. Here, we describe the docking of 2, 4-diamino-5-methyl-5-deazapteridine (DMDP) derivatives as inhibitors to human dihydrofolate reductase (DHFR). We docked 78 DMDP derivates collected from literature to DHFR and studied their specific interactions with DHFR. A new shape-based method, LigandFit, was used for docking DMDP derivatives into DHFR active sites. The result indicates that the molecular docking approach is reliable and produces a good correlation coefficient (r2 = 0.499) for the 73 compounds between docking score and IC50 values (Inhibitory Activity). The chloro substituted naphthyl ring of compound 63 makes significant hydrophobic contact with Leu 22, Phe 31 and Pro 61 of the DHFR active site leading to enhanced inhibition of the enzyme. The docked complexes provide better insights to design more potent DHFR inhibitors prior to their synthesis.  相似文献   

5.
Cody V  Galitsky N  Rak D  Luft JR  Pangborn W  Queener SF 《Biochemistry》1999,38(14):4303-4312
Structural data from two independent crystal forms (P212121 and P21) of the folate (FA) binary complex and from the ternary complex with the oxidized coenzyme, NADP+, and recombinant Pneumocystis carinii dihydrofolate reductase (pcDHFR) refined to an average of 2.15 A resolution, show the first evidence of ligand-induced conformational changes in the structure of pcDHFR. These data are also compared with the crystal structure of the ternary complex of methotrexate (MTX) with NADPH and pcDHFR in the monoclinic lattice with data to 2.5 A resolution. Comparison of the data for the FA binary complex of pcDHFR with those for the ternary structures reveals significant differences, with a >7 A movement of the loop region near residue 23 that results in a new "flap-open" position for the binary complex, and a "closed" position in the ternary complexes, similar to that reported for Escherichia coli (ec) DHFR complexes. In the orthorhombic lattice for the binary FA pcDHFR complex, there is also an unwinding of a short helical region near residue 47 that places hydrophobic residues Phe-46 and Phe-49 toward the outer surface, a conformation that is stabilized by intermolecular packing contacts. The pyrophosphate moiety of NADP+ in the ternary folate pcDHFR complexes shows significant differences in conformation compared with that observed in the MTX-NADPH-pcDHFR ternary complex. Additionally, comparison of the conformations among these four pcDHFR structures reveals evidence for subdomain movement that correlates with cofactor binding states. The larger binding site access in the new "flap-open" loop 23 conformation of the binary FA complex is consistent with the rapid release of cofactor from the product complex during catalysis as well as the more rapid release of substrate product from the binary complex as a result of the weaker contacts of the closed loop 23 conformation, compared to ecDHFR.  相似文献   

6.
The structure of rat intestinal fatty acid binding protein (I-FABP) with bound oleate (C18:1) has been refined with x-ray diffraction data to a resolution of 1.75 A. The protein contains 10 anti-parallel beta strands composed of 99 residues and 2 short helices of 14 residues. Oleate is located in the interior of the protein in a bent conformation with C1-C12 more ordered than C13-C18. Two of the eight ordered waters in I-FABP:oleate are part of a hydrogen bond network that includes the carboxylate of oleate, the guanidinium group of Arg106, the nitrogen of the indole group of Trp82, and the side chain of Gln115. Most of the methylenes of bound oleate reside in a crevice formed by hydrophobic and aromatic side chains. Tyr70 and Tyr117 envelop the acyl chain from C3 to C8 forming contacts with both the convex and concave faces of its van der Waals surface. The hydroxyls of each phenolic side chain hydrogen bond to ordered water molecules. Two ordered waters make van der Waals contact with the concave face of the bound fatty acid. The omega-terminal methyl of oleate is oriented so that it points toward the center of the benzene of Phe55 allowing it to form van der Waals interactions with its component methylenes. Comparison of the structure of I-FABP:oleate with a recently refined 1.19-A model of apoI-FABP and an earlier 2.0-A model of I-FABP:palmitate revealed a remarkable degree of similarity in the positions of their main chain and side chain atoms and in the conformations of the bound oleate and palmitate. The principal differences were confined to a few discrete regions of the protein. The helical domain, the type I turn between beta strands C and D, and the ring of Phe55 together form a solvent-accessible portal to the interior of the protein. They are repositioned in I-FABP:oleate (and I-FABP:palmitate) so that the binding cavity is even more accessible to solvent and its volume is increased. The side chain of Phe55 which shows discrete disorder in the apoprotein functions as an omega-terminal "sensing device": moving progressively outward toward the surface as the chain length of the bound fatty acid increases by 2 methylenes. Tyr70 and Tyr117 which also show discrete disorder in the apoprotein structure due to rotation around their C alpha-C beta bonds, are stabilized in a single, well ordered position in the holoproteins.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
A theoretical conformational analysis (molecular mechanics study) of nine cyclic tetrapeptides, structurally related to the highly mu-receptor-selective dermorphin analogue H-Tyr-D-Orn-Phe-Asp-NH2, was performed. These compounds display considerable diversity in their mu-receptor affinity and selectivity. A systematic search and subsequent energy minimization in absence of the exocyclic Tyr1 residue and Phe3 side chain revealed the constrained nature of the 11-13-membered ring structures contained in these analogues. No more than four low-energy conformers (within 2 kcal/mol of the lowest energy conformation) were found in each case. After attachment of the Tyr1 moiety and Phe3 side chain to the "bare" low-energy ring structures, a systematic search and energy minimization of these exocyclic moieties resulted in a limited number of low-energy conformational families for all compounds. Five analogues with high mu-receptor affinity--H-Tyr-D-Orn-Phe-Asp-NH2, H-Tyr-D-Orn-Phe-D-Asp-NH2, H-Tyr-D-Asp-Phe-Orn-NH2, H-Tyr-D-Asp-Phe-A2bu-NH2 (A2 bu: alpha, gamma-diaminobutyric acid) and H-Tyr-D-Cys-Phe-Cys-NH2--all showed a tilted stacking interaction between the Tyr1 and Phe3 aromatic rings in the lowest or second lowest energy conformation found. The same kind of stacking was not possible in low-energy conformers of the four analogues with poor affinity for the mu-receptor [H-Tyr-L-Orn-Phe-Asp-NH2, H-Tyr-D-Orn-D-Phe-Asp-NH2, H-Tyr-D-Orn-Phe(NMe)-Asp-NH2 [Phe(NMe): N alpha-methylphenylalanine], and H-Tyr-D-Orn-Phg-Asp-NH2 (Phg: phenylglycine)].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The crystal structures of two human dihydrofolate reductase (hDHFR) ternary complexes, each with bound NADPH cofactor and a lipophilic antifolate inhibitor, have been determined at atomic resolution. The potent inhibitors 6-([5-quinolylamino]methyl)-2,4-diamino-5-methylpyrido[2,3-d]pyrimidine (SRI-9439) and (Z)-6-(2-[2,5-dimethoxyphenyl]ethen-1-yl)-2,4-diamino-5-methylpyrido[2,3-d]pyrimidine (SRI-9662) were developed at Southern Research Institute against Toxoplasma gondii DHFR-thymidylate synthase. The 5-deazapteridine ring of each inhibitor adopts an unusual puckered conformation that enables the formation of identical contacts in the active site. Conversely, the quinoline and dimethoxybenzene moieties exhibit distinct binding characteristics that account for the differences in inhibitory activity. In both structures, a salt-bridge is formed between Arg70 in the active site and Glu44 from a symmetry-related molecule in the crystal lattice that mimics the binding of methotrexate to DHFR.  相似文献   

9.
The stability constants of a series of complexes of the cationic water-soluble porphyrin ZnTMPyP with various amino acids have been determined by 1H NMR spectroscopy at pH 10.5. The following stability order has been observed: Tyr greater than Phe, Glu greater than Asp greater than Ile greater than Val greater than Gly. These results can be best rationalized by invoking complex stabilization due to ligand-ligand (e.g., stacking or electrostatic) interactions. Evidence for stacking interactions between the porphyrin ring and the aromatic ring of phenylalanine, tyrosine, and tryptophan was further provided by study of the complexation of these amino acids with the free-base porphyrin TMPyPH2. In this case, complexation constants increased in the order: Phe less than Tyr less than Trp. Attempts to form complexes of the amino acids with the anionic porphyrin ZnTCPP proved unsuccessful, indicating that electrostatic interactions play a major role in the stability of the zinc porphyrin-amino acids complexes.  相似文献   

10.
2,4-Diamino-5-[3',4'-dimethoxy-5'-(5-carboxy-1-pentynyl)]benzylpyrimidine (6) and 2,4-diamino-5-[3',4'-dimethoxy-5'-(4-carboxyphenylethynyl)benzylpyrimidine (7) were synthesized from 2,4-diamino-5-(5'-iodo-3',4'-dimethoxybenzyl)pyrimidine (9) via a Sonogashira reaction with appropriate acetylenic esters followed by saponification, and were tested as inhibitors of dihydrofolate reductase (DHFR) from Pneumocystis carinii (Pc), Toxoplasma gondii (Tg), Mycobacterium avium (Ma), and rat in comparison with the widely used antibacterial agent 2,4-diamino-5-(3',4',5'-trimethoxybenzyl)pyrimidine (trimethoprim, TMP). The selectivity index (SI) for each compound was calculated by dividing its 50% inhibitory concentration (IC(50)) against rat DHFR by its IC(50) against Pc, Tg, or Ma DHFR. The IC(50) of 6 against Pc DHFR was 1.0 nM, with an SI of 5000. Compound 7 had an IC(50) of 8.2 nM against Ma DHFR, with an SI of 11000. By comparison, the IC(50) of TMP was 12000 nM against Pc, 300 nM against Ma, and 180000 against rat DHFR. The potency and selectivity values of 6 and 7 were not as high against Tg as they were against Pc or Ma DHFR, but nonetheless exceeded those of TMP. Because of the outstanding selectivity of 6 against Pc and of 7 against Ma DHFR, these novel analogues may be viewed as promising leads for further structure-activity optimization.  相似文献   

11.
Coetsee M  Millar RP  Flanagan CA  Lu ZL 《Biochemistry》2008,47(39):10305-10313
Molecular modeling showed interactions of Tyr (290(6.58)) in transmembrane domain 6 of the GnRH receptor with Tyr (5) of GnRH I, and His (5) of GnRH II. The wild-type receptor exhibited high affinity for [Phe (5)]GnRH I and [Tyr (5)]GnRH II, but 127- and 177-fold decreased affinity for [Ala (5)]GnRH I and [Ala (5)]GnRH II, indicating that the aromatic ring in position 5 is crucial for receptor binding. The receptor mutation Y290F decreased affinity for GnRH I, [Phe (5)]GnRH I, GnRH II and [Tyr (5)]GnRH II, while Y290A and Y290L caused larger decreases, suggesting that both the para-OH and aromatic ring of Tyr (290(6.58)) are important for binding of ligands with aromatic residues in position 5. Mutating Tyr (290(6.58)) to Gln increased affinity for Tyr (5)-containing GnRH analogues 3-12-fold compared with the Y290A and Y290L mutants, suggesting a hydrogen-bond between Gln of the Y290Q mutant and Tyr (5) of GnRH analogues. All mutations had small effects on affinity of GnRH analogues that lack an aromatic residue in position 5. These results support direct interactions of the Tyr (290(6.58)) side chain with Tyr (5) of GnRH I and His (5) of GnRH II. Tyr (290(6.58)) mutations, except for Y290F, caused larger decreases in GnRH potency than affinity, indicating that an aromatic ring is important for the agonist-induced receptor conformational switch.  相似文献   

12.
A concise new route allowing easy access to five previously unreported 2,4-diamino-6-(substituted benzyl)pyrido[2,3-d]pyrimidines (2a-e) was developed, involving condensation of 2,4-dipivaloylamino-5-bromopyrido[2,3-d]pyrimidine (6) with an organozinc halide in the presence of a catalytic amount of [1,1'-bis(diphenylphosphino)ferrocene]dichloropalladium(II).CH(2)Cl(2), followed by removal of the pivaloyl groups with base. Also prepared via a scheme based on the Taylor ring expansion/ring annulation synthesis were three heretofore undescribed 2,4-diamino-5-(substituted benzyl)-7H-pyrrolo[2,3-d]pyrimidines (3b-c). Standard spectrophotometric assays were used to compare the ability of 2a-e and 3b-c to inhibit dihydrofolate reductase (DHFR) from Pneumocystis carinii, Toxoplasma gondii, and Mycobacterium avium, three examples of opportunistic pathogens to which AIDS patients are highly vulnerable because of their immunocompromised state. For comparison, 13 previously untested 2,4-diamino-6-(substituted benzyl)quinazolines (17a-m) were also evaluated as inhibitors of these enzymes, as well as the enzyme from rat liver. None of the quinazolines or pyridopyrimidines tested was more potent against the P. carinii enzyme than the structurally related reference compound piritrexim (1), and none showed selectivity for the P. carinii enzyme over the rat enzyme. One of the pyridopyrimidines (2c) showed 10-fold selectivity for T. gondii versus rat DHFR, and two of them (2b, 2c) showed selectivity for the M. avium enzyme. However, this gain in species selectivity was achieved at the cost of decreased in potency, as has been noted with many other lipophilic DHFR inhibitors.  相似文献   

13.
Aromatic amino acid aminotransferase (ArATPh), which has a melting temperature of 120 degrees C, is one of the most thermostable aminotransferases yet to be discovered. The crystal structure of this aminotransferase from the hyperthermophilic archaeon Pyrococcus horikoshii was determined to a resolution of 2.1 A. ArATPh has a homodimer structure in which each subunit is composed of two domains, in a manner similar to other well characterized aminotransferases. By the least square fit after superposing on a mesophilic ArAT, the ArATPh molecule exhibits a large deviation of the main chain coordinates, three shortened alpha-helices, an elongated loop connecting two domains, and a long loop transformed from an alpha-helix, which are all factors that are likely to contribute to its hyperthermostability. The pyridine ring of the cofactor pyridoxal 5'-phosphate covalently binding to Lys(233) is stacked parallel to F121 on one side and interacts with the geminal dimethyl-CH/pi groups of Val(201) on the other side. This tight stacking against the pyridine ring probably contributes to the hyperthermostability of ArATPh. Compared with other ArATs, ArATPh has a novel substrate specificity, the order of preference being Tyr > Phe > Glu > Trp > His> Met > Leu > Asp > Asn. Its relatively weak activity against Asp is due to lack of an arginine residue corresponding to Arg(292)* (where the asterisk indicates that this is a residues supplied by the other subunit of the dimer) in pig cytosolic aspartate aminotransferase. The enzyme recognizes the aromatic substrate by hydrophobic interaction with aromatic rings (Phe(121) and Tyr(59)*) and probably recognizes acidic substrates by a hydrophilic interaction involving a hydrogen bond network with Thr(264)*.  相似文献   

14.
A detailed analysis of the composition and properties of hydrophobic nuclei and microclusters in pancreatic ribonuclease A (RNase A) has been carried out. Distance calculations for all noncovalently bonded atoms revealed that the average number of nonpolar contacts between a side chain of an amino acid and its neighbors is substantially larger if it involves hydrophobic residues rather than nonhydrophobic ones. However, the difference decreased when the number of contacts per nonpolar group and/or atom were calculated. Three main nuclei and five microclusters were identified, and their quantitative parameters were calculated. These nuclei include hydrophobic residues with a substantial number of nonpolar contacts with the environment (Phe 8, Phe 120, Phe 46, Tyr 25, Tyr 97, Ile 107, Leu 35, Ile 81, Val 54, Val 108, Met 29, Met 30). Hydrophobic nuclei of RNase A differ in shape and in composition, in the number of intranuclear contacts and of associated residues, as well as in their internal mobility. All eight cysteine residues are involved in nonpolar interactions with amino acid residues of hydrophobic nuclei. Active site amino acid residues of RNase A form a noncovalent contact network comprised of themselves, as well as of many conserved residues from hydrophobic nuclei. Sequence alignment with some other members of the RNase A family of proteins shows remarkable similarity in positions and in conservation of the main nonpolar residues, comprising cores of two (out of three) hydrophobic nuclei. A correlation was shown to exist between the average density of contacts for side-chain atoms and the number of amino acids to be found in the appropriate positions in the sequences of related mammalian ribonucleases. However, there are certain amino acid positions in the third, smaller nucleus, which are highly variable within the family. Taking into account that this nucleus is composed of residues belonging to different elements of the secondary structure, it is likely that the mutual orientation of these elements can be somehow different for these proteins.  相似文献   

15.
The stacking interactions of the clonidine aromatic ring with the aromatic rings of Phe or Tyr of alpha2-adrenoreceptor and Tyr aromatic ring of the pore of the tetradotoxin-resistant channel have been investigated. Ab initio quantum chemical calculations for a model system of two parallel aromatic rings were performed by GAMESS software with 6-31G** basis set in the framework of the Moller-Plesset second-order perturbation theory with full geometry optimization without any symmetry. It was shown that the parallel shifted conformation of two aromatic rings is energetically most favorable. The 2,6-chlorination of one of the benzene rings leads to the amplification of the stacking interaction, an increase in the relative shift of the rings and possible growth of both the hypotensive and analgetic functions of clonidine due to the increase in the binding energy. The 4-fluoridization of the clonidine benzene ring can amplify its analgetic function but practically excludes its hypotensive action.  相似文献   

16.
Synthesis of 2,4-diamino-5-cyano-6-[[(diisopropoxyphosphoryl)methoxy]ethoxy]pyrimidine was based on the formation of the pyrimidine ring by cyclization followed by modification of the side chain by alkylation. The 5-cyano group was also transformed to a 5-formyl and 5-hydroxymethyl group by reduction. As a side product an unexpected dimer was formed. Resulting compounds were converted to the free phosphonic acids by treatment with bromotrimethylsilane followed by hydrolysis. The 5-cyano and 5-formyl derivatives showed pronounced antiretroviral activity, comparable to that of the reference drugs adefovir and tenofovir.  相似文献   

17.
The high-resolution 3D structure of the octapeptide hormone angiotensin II (AII) in aqueous solution has been obtained by simulated annealing calculations, using high-resolution NMR-derived restraints. After final refinement in explicit water, a family of 13 structures was obtained with a backbone RMSD of 0.73 +/- 0.23 A. AII adopts a fairly compact folded structure, with its C-terminus and N-terminus approaching to within approximately 7.2 A of each other. The side chains of Arg2, Tyr4, Ile5 and His6 are oriented on one side of a plane defined by the peptide backbone, and the Val3 and Pro7 are pointing in opposite directions. The stabilization of the folded conformation can be explained by the stacking of the Val3 side chain with the Pro7 ring and by a hydrophobic cluster formed by the Tyr4, Ile5 and His6 side chains. Comparison between the NMR-derived structure of AII in aqueous solution and the refined crystal structure of the complex of AII with a high-affinity mAb (Fab131) [Garcia, K.C., Ronco, P.M., Verroust, P.J., Brunger, A.T., Amzel, L.M. (1992) Science257, 502-507] provides important quantitative information on two common structural features: (a) a U-shaped structure of the Tyr4-Ile5-His6-Pro7 sequence, which is the most immunogenic epitope of the peptide, with the Asp1 side chain oriented towards the interior of the turn approaching the C-terminus; (b) an Asx-turn-like motif with the side chain aspartate carboxyl group hydrogen-bonded to the main chain NH group of Arg2. It can be concluded that small rearrangements of the epitope 4-7 in the solution structure of AII are required by a mean value of 0.76 +/- 0.03 A for structure alignment and approximately 1.27 +/- 0.02 A for sequence alignment with the X-ray structure of AII bound to the mAb Fab131. These data are interpreted in terms of a biological "nucleus" conformation of the hormone in solution, which requires a limited number of structural rearrangements for receptor-antigen recognition and binding.  相似文献   

18.
In order to produce a more potent replacement for trimethoprim (TMP) used as a therapy for Pneumocystis pneumonia and targets dihydrofolate reductase from Pneumocystis jirovecii (pjDHFR), it is necessary to understand the determinants of potency and selectivity against DHFR from the mammalian host and fungal pathogen cells. To this end, active site residues in human (h) DHFR were replaced with those from pjDHFR. Structural data are reported for two complexes of TMP with the double mutants Gln35Ser/Asn64Phe (Q35S/N64F) and Gln35Lys/Asn64Phe (Q35K/N64F) of hDHFR that unexpectedly show evidence for the binding of two molecules of TMP: one molecule that binds in the normal folate binding site and the second molecule that binds in a novel subpocket site such that the mutated residue Phe64 is involved in van der Waals contacts to the trimethoxyphenyl ring of the second TMP molecule. Kinetic data for the binding of TMP to hDHFR and pjDHFR reveal an 84-fold selectivity of TMP against pjDHFR (Ki 49 nM) compared to hDHFR (Ki 4093 nM). Two mutants that contain one substitution from pj- and one from the closely related Pneumocystis carinii DHFR (pcDHFR) (Q35K/N64F and Q35S/N64F) show Ki values of 593 and 617 nM, respectively; these Ki values are well above both the Ki for pjDHFR and are similar to pcDHFR (Q35K/N64F and Q35S/N64F) (305 nM). These results suggest that active site residues 35 and 64 play key roles in determining selectivity for pneumocystis DHFR, but that other residues contribute to the unique binding of inhibitors to these enzymes.  相似文献   

19.
The three-dimensional X-ray structure of the RNase T1[EC 3.1.27.3]-2'GMP complex crystallized at low pH value (4.0) was determined, and refined to 1.9 A resolution to give a final R value of 0.203. The refined model includes 781 protein atoms, 24 inhibitor atoms, and 43 solvent molecules. The imidazole rings of His27 and His40 interact with the carboxyl side chains of Glu82 and Glu58, respectively, whereas that of His92 is in contact with the main chain carbonyl oxygen of Ala75. In the complex, the ribose ring of the 2'GMP molecule adopts a C2'-endo puckering, and the exocyclic conformation is gauche(-)-gauche(+). The glycosyl torsion angle is in the syn range with an intramolecular hydrogen bond between N3 and O5', and the 2'-phosphate orientation is trans-gauche(-). The guanine base of the inhibitor is tightly bound to the base recognition site with five hydrogen bonds (N1--Glu46O epsilon 2, N2---Asn98O,O6---Asn44N, and N7 ---Asn43N delta 2/Asn43N) and is sandwiched between the phenolic ring portions of Tyr42 and Tyr45 by stacking interactions. The 2'-phosphate group interacts with Arg77N eta 2, Glu58O episilon 2, and Tyr 38O eta but not with any of the histidine residues. Arg77N eta 2 also interacts with Tyr38O eta. There is no interaction between the ribose moiety of the inhibitor and the enzyme.  相似文献   

20.
Stacking interactions of the clonidine aromatic ring with that of Phe or Tyr in the α2-adrenoreceptor and Tyr in the tetrodotoxin-resistant sodium channel pore have been studied. Ab initio quantum-chemical calculations for a model system of two parallel aromatic rings have been performed with the GAMESS software using the 6-31G* basis set in the framework of the second-order Muller-Plesset perturbation theory with full geometry optimization without symmetry constraints. The parallel shifted conformation of two aromatic rings was found to be energetically most favorable. The 2′,6′-chlorination of one of the benzene rings enhances the stacking interaction, somewhat increases the shift of these rings, and possibly improves the hypotensive and analgesic functions of clonidine owing to an increase in the binding energy. The 4-fluorination of the clonidine ring can increase its analgesic effect but practically excludes its hypotensive activity.  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