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 phenolic side chain of Tyr(4) present in Ang II is proposed to interact with the side chain of Arg 167 of the AT1 receptor. To determine the contribution of the analogous Arg182 in the ligand-binding properties of the AT2, we replaced the Arg182 with Glu and Ala, and analyzed the ligand-binding properties. Our results suggest that replacing Arg182 with either Glu or Ala abolished the ability of the AT2 receptor to bind the nonspecific peptidic ligands, (125)I-Ang II and [(125)I-Sar(1)-Ile(8)]Ang II, as well as the AT2 receptor-specific peptidic ligand (125)I-CGP42112A. We have shown previously that replacing the positively charged side chain of Lys215 with the negatively charged side chain of Glu in the fifth TMD did not alter the high affinity binding of (125)I-CGP42112A to the AT2 receptor. However, ligand-binding properties of the Arg182Glu mutant suggest that positively charged side chain of Arg182 located in the junction of second ECL and the fourth TMD is critical for high affinity binding of all three peptidic ligands to the AT2 receptor.  相似文献   

2.
In order to seek vancomycin analogs with improved performance against VanA and VanB resistant bacterial strains, extensive computational investigations have been performed to examine the effects of side-chain and backbone modifications. Changes in binding affinities for tripeptide cell-wall precursor mimics, Ac2-l-Lys-d-Ala-d-Ala (3) and Ac2-l-Lys-d-Ala-d-Lac (4), with vancomycin analogs were computed with Monte Carlo/free energy perturbation (MC/FEP) calculations. Replacements of the 3-hydroxyl group in residue 7 with small alkyl or alkoxy groups, which improve contacts with the methyl side chain of the ligands’ d-Ala residue, are predicted to be the most promising to enhance binding for both ligands. The previously reported amine backbone modification as in 5 is shown to complement the hydrophobic modifications for binding monoacetylated tripeptides. In addition, replacement of the hydroxyl groups in residues 5 and 7 by fluorine is computed to have negligible impact on binding the tripeptides, though it may be pharmacologically advantageous.  相似文献   

3.
Tjong SC  Chen TS  Huang WN  Wu WG 《Biochemistry》2007,46(35):9941-9952
Cobra cardiotoxins (CTXs) are three-fingered polypeptides with positively charged domains that have been shown to bind to anionic ligands of snake venom citrate, glycosaminoglycans, sulfoglycosphingolipid, and nucleotide triphosphate with various biochemical effects including toxin dimerization, cell surface retention, membrane pore formation, cell internalization and blocking of enzymatic activities of kinase and ATPase. The reported anionic binding sites, however, are found to be different among different CTX homologues for potentially different CTX activities. Herein, by NMR studies of the binding of inorganic phosphate, dATP (stable form of ATP), and heparin-derived tetrasaccharide to Naja atra CTX A1, a novel CTX molecule exhibiting in vivo necrotic activity on skeletal muscle, we demonstrate that diverse ligands binding to CTXs could also occur at a single protein site with flexible side chain interactions. The flexibility of such an interaction is also illustrated by the available heparin-CTX A3 complex structures with different heparin chain lengths binding at the same site. Our results provide a likely structural explanation on how the interaction between heparan sufate and proteins depends more on the overall charge cluster organization rather than on their fine structures. We also suggest that the ligand binding site of CTX homologues can be fine-tuned by nonconserved residues near the binding pocket because of their flexible side chain interaction and dimerization ability, even for the rigid CTX molecules tightened by four disulfide bonds.  相似文献   

4.
Nucleophile specificity of subtilisin (subtilopeptidase A) was studied via acyl transfer reactions in acetonitrile containing piperidine and 10 vol% of water. Ac-Tyr-OEt was used as acyl donor and a series of amino acid derivatives, di- and tripeptides of the general structure Xaa-Gly, Gly-Xaa, Gly-Gly-Xaa (Xaa represents all natural L-amino acids except cysteine) were used as nucleophiles. The nucleophilic efficiencies of these peptides were characterized by the values of the apparent partition constants, p(app), determined from the HPLC analysis of the reactions. The order of preference for the P'(1) position was estimated to be: Gly > hydrophilic, positively charged > hydrophobic, aromatic > negatively charged > Leu > Pro side chain. For the P'(2) position the order of preference was: Gly > hydrophilic, charged > hydrophobic, aromatic > Pro side chain. The values of p(app) for Gly-Gly-Xaa tripeptides cover a range of only two orders of magnitude, with lower nucleophile efficiency for those with hydrophobic amino acid residues in the P'(3) position. The dipeptide with Pro in P'(1) did not react at all, but a tripeptide having Pro in P'(3) was a very good nucleophile. The negatively charged amino acid residues in the P'(1) position result in very weak nucleophilic behavior, whereas the peptides with Asp or Glu in P'(2) and P'(3) are well accepted. Generally, peptides of the Gly-Xaa or Gly-Gly-Xaa series were better nucleophiles than peptides of the Xaa-Gly series. The length of the peptide chain or amidation of alpha-carboxyl function had no influence on nucleophilic behavior. No significant difference in nucleophile specificity between subtilopeptidase A and nagarse was observed. (c) 1996 John Wiley & Sons, Inc.  相似文献   

5.
A transport double mutant of Neurospora crassa has been isolated that has only one of the three transport systems capable of l-histidine uptake. The substrate specificity of the remaining transport system, termed the general transport system, has been fully characterized with regard to the contributions to binding of the side chain, the alpha-amino group, and the carboxylate group. The positively charged alpha-amino group is necessary for binding; the negatively charged carboxylate group is of less importance, since its replacement by a neutral carbonyl functional group does not completely abolish binding. The greatest structural latitude for binding was found in the side chain; affinity for alpha-amino acids was uniformly high except for l-aspartic and l-glutamic acids, l-asparagine, and l-proline. Thus, this transport system is "general" with these restrictions.  相似文献   

6.
The interactions of a homologous series of four anthraquinone (AQ) intercalators with increasing lengths of polyethylene glycol (PEG) side chains with DNA have been studied via molecular dynamics (MD) simulations. The geometry, conformation, interactions, and hydration of the complexes were examined. The geometries of the four ligands were similar with parallel stacking to the long axis of the adjacent DNA base pairs. Hydrogen bonding between the AQ amide and DNA led to a preference for the trans-syn conformer. As the side chain lengthened, binding to DNA reduced the conformational space, resulting in an increase in unfavorable entropy. Increased localization of the PEG side chain in the DNA groove, indicating some interaction of the side chain with DNA, also contributed unfavorably to the entropy. The changes in free energy of binding due to entropic considerations (-3.9 to -6.3 kcal/mol) of AQ I-IV were significant. The hydration of the PEG side chain decreased upon binding to DNA. Understanding of side chain conformations, interactions, and hydration changes that accompany the formation of a ligand-DNA complex may be important in the development of new applications of pegylated small molecules that target biological macromolecules.  相似文献   

7.
Binding of ionic ligands to polyelectrolytes.   总被引:1,自引:1,他引:0       下载免费PDF全文
Ionic ligands can bind to polyelectrolytes such as DNA or charged polysaccharides. We develop a Poisson-Boltzmann treatment to compute binding constants as a function of ligand charge and salt concentration in the limit of low ligand concentration. For flexible chain ligands, such as oligopeptides, we treat their conformations using lattice statistics. The theory predicts the salt dependence and binding free energies, of Mg(2+) ions to polynucleotides, of hexamine cobalt(III) to calf thymus DNA, of polyamines to T7 DNA, of oligolysines to poly(U) and poly(a), and of tripeptides to heparin, a charged polysaccharide. One parameter is required to obtain absolute binding constants, the distance of closest separation of the ligand to the polyion. Some, but not all, of the binding entropies and enthalpies are also predicted accurately by the model.  相似文献   

8.
Abstract

The interactions of a homologous series of four anthraquinone (AQ) intercalators with increasing lengths of polyethylene glycol (PEG) side chains with DNA have been studied via molecular dynamics (MD) simulations. The geometry, conformation, interactions, and hydration of the complexes were examined. The geometries of the four ligands were similar with parallel stacking to the long axis of the adjacent DNA base pairs. Hydrogen bonding between the AQ amide and DNA led to a preference for the trans-syn conformer. As the side chain lengthened, binding to DNA reduced the conformational space, resulting in an increase in unfavorable entropy. Increased localization of the PEG side chain in the DNA groove, indicating some interaction of the side chain with DNA, also contributed unfavorably to the entropy. The changes in free energy of binding due to entropic considerations (— 3.9 to—6.3 kcal/mol) of AQ I-IV were significant. The hydration of the PEG side chain decreased upon binding to DNA. Understanding of side chain conformations, interactions, and hydration changes that accompany the formation of a ligand-DNA complex may be important in the development of new applications of pegylated small molecules that target biological macromolecules.  相似文献   

9.
Conformational studies were performed on a synthetic pentacosapeptide representing the RNA-binding N-terminal region of the coat protein of cowpea chlorotic mottle virus. The effects of ionic strength, addition of (oligo)phosphates and temperature on the conformation of this highly positively charged peptide containing six arginines and three lysines were studied. CD experiments show that the peptide has 15-18% alpha-helical conformation and about 80% random-coil conformation in the absence of inorganic salt at 25 degrees C, and 20-21% alpha-helical conformation under the same conditions at 10 degrees C. Addition of inorganic salts results in an increase of alpha-helix content, up to 42% in the presence of oligophosphate with an average chain length of 18 phosphates, which was used as an RNA analog. NMR experiments show that the alpha-helix formation starts in the region between Thr9 and Gln12, and is extended in the direction of the C terminus. Relaxation measurements show that binding to oligophosphates of increasing length results in reduced internal mobilities of the positively charged side chains of the arginyl and lysyl residues and of the side chain of Thr9 in the alpha-helical region. The alpha-helix formation in the N-terminal part of this viral coat protein upon binding of phosphate groups to the positively charged side chains is suggested to play an essential role in RNA binding.  相似文献   

10.
Telomerase is responsible for the immortal phenotype of cancer cells and telomerase inhibition may specifically target cancer cell proliferation. Ligands able to selectively bind to G-quadruplex telomeric DNA have been considered as telomerase inhibitors but their mechanisms of action have often been deduced from a non-quantitative telomerase activity assay (TRAP assay) that involves a PCR step and that does not provide insight on the mechanism of inhibition. Furthermore, quadruplex ligands have also been shown to exert their effects by affecting association of telomere binding proteins with telomeres. Here, we use quantitative direct telomerase activity assays to evaluate the strength and mechanism of action of hydrosoluble perylene diimides (HPDIs). HPDIs contain a perylene moiety and different numbers of positively charged side chains. Side chain features vary with regard to number and distances of the charges. IC50 values of HPDIs were in the low micromolar (0.5–5 μM) range depending on the number and features of the side chains. HPDIs having four side chains emerged as the best compounds of this series. Analysis of primer elongation products demonstrated that at low HPDI concentrations, telomerase inhibition involved formation of telomeric G-quadruplex structures, which inhibited further elongation by telomerase. At high HPDI concentrations, telomerase inhibition occurred independently of G-quadruplex formation of the substrate. The mechanism of action of HPDIs and their specific binding to G-quadruplex DNA was supported by PAGE analysis, CD spectroscopy and ESI-MS. Finally, competition Telospot experiments with duplex DNA indicated specific binding of HPDIs to the single-stranded telomeric substrates over double stranded DNA, a result supported by competitive ESI-MS. Altogether, our results indicate that HPDIs act by stabilizing G-quadruplex structures in single-stranded telomeric DNA, which in turn prevents repeat addition processivity of telomerase.  相似文献   

11.
We have determined the apparent molar volumes and the apparent molar adiabatic compressibilities at 25 degrees C of 10 X-Gly-Gly and Gly-Gly-X tripeptides in which X represents a residue with a non-polar side chain. We also have determined the changes in volume and compressibility which accompany neutralization of the amino and carboxyl termini in these tripeptides. The mutual influence of the non-polar side chain of the X residue and the terminal amino and carboxyl groups on the hydration of each other depends on the chemical nature of the side chain and the state of ionization of the termini. We interpret our data in terms of the hydration of the component aliphatic, aromatic, and charged atomic groups, as well as the mutual interactions between these groups.  相似文献   

12.
The Escherichia coli peptide binding protein OppA is an essential component of the oligopeptide transporter Opp. Based on studies on its orthologue from Salmonella typhimurium, it has been proposed that OppA binds peptides between two and five amino acids long, with no apparent sequence selectivity. Here, we studied peptide binding to E. coli OppA directly and show that the protein has an unexpected preference for basic peptides. OppA was expressed in the periplasm, where it bound to available peptides. The protein was purified in complex with tightly bound peptides. The crystal structure (up to 2.0 Å) of OppA liganded with the peptides indicated that the protein has a preference for peptides containing a lysine. Mass spectrometry analysis of the bound peptides showed that peptides between two and five amino acids long bind to the protein and indeed hinted at a preference for positively charged peptides. The preference of OppA for peptides with basic residues, in particular lysines, was corroborated by binding studies with peptides of defined sequence using isothermal titration calorimetry and intrinsic protein fluorescence titration. The protein bound tripeptides and tetrapeptides containing positively charged residues with high affinity, whereas related peptides without lysines/arginines were bound with low affinity. A structure of OppA in an open conformation in the absence of ligands was also determined to 2.0 Å, revealing that the initial binding site displays a negative surface charge, consistent with the observed preference for positively charged peptides. Taken together, E. coli OppA appears to have a preference for basic peptides.  相似文献   

13.
The thermodynamics of binding of unfolded polypeptides to the chaperone SecB was investigated in vitro by isothermal titration calorimetry and fluorescence spectroscopy. The substrates were reduced and carboxamidomethylated forms of RNase A, BPTI, and alpha-lactalbumin. SecB binds both fully unfolded RNase A and BPTI as well as compact, partially folded disulfide intermediates of alpha-lactalbumin, which have 40-60% of native secondary structure. The heat capacity changes observed on binding the reduced and carboxamidomethylated forms of alpha-lactalbumin, BPTI, and RNase A were found to be -0.10, -0.29, and -0.41 kcal mol(-1) K(-1), respectively, and suggest that between 7 and 29 residues are buried upon substrate binding to SecB. In all cases, binding occurs with a stoichiometry of one polypeptide chain per monomer of SecB. There is no evidence for two separate types of binding sites for positively charged and hydrophobic ligands. Spectroscopic and proteolysis protection studies of the binding of SecB to poly-L-Lys show that binding of highly positively charged peptide ligands to negatively charged SecB leads to charge neutralization and subsequent aggregation of SecB. The data are consistent with a model where SecB binds substrate molecules at an exposed hydrophobic cleft. SecB aggregation in the absence of substrate is prevented by electrostatic repulsion between negatively charged SecB tetramers.  相似文献   

14.
Summary One Lys/Phe copolymer and two series of copolymers of lysine with either alanine or tyrosine have been used as inhibitors of a plant proteinase that is known to be inhibited by polycationic inhibitors. The copolymers differ in the hydrophobicity of the non-lysine amino acid residue and the single members of each series differ from each other in their degree of polymerization and in their charge density, i.e., the frequency of occurrence of the lysine residue in the synthetic polyamino acid chain. All the tested copolymers show cooperative inhibition, with a Hill coefficient higher than 1. CD measurements indicate that the inhibition is realized through a conformational change of the enzyme molecule. Both the enzyme inhibition and the conformational change are supported by aspecific electrostatic binding between the positively charged groups of the lysine moiety and the negatively charged groups of the enzyme surface. In each series the inhibitory power increases with the charge density, while at the same charge density the inhibitory efficiency depends on the hydrophobicity of the side chain of the non-lysine amino acid in the order Phe>Tyr>Ala.  相似文献   

15.
Electrostatic effects play a central role in cold adaptation of trypsin.   总被引:2,自引:0,他引:2  
Organisms that live in constantly cold environments have to adapt their metabolism to low temperatures, but mechanisms of enzymatic adaptation to cold environments are not fully understood. Cold active trypsin catalyses reactions more efficiently and binds ligands more strongly in comparison to warm active trypsin. We have addressed this issue by means of comparative free energy calculations studying the binding of positively charged ligands to two trypsin homologues. Stronger inhibition of the cold active trypsin by benzamidine and positively charged P1-variants of BPTI is caused by rather subtle electrostatic effects. The different affinity of benzamidine originates solely from long range interactions, while the increased binding of P1-Lys and -Arg variants of BPTI is attributed to both long and short range effects that are enhanced in the cold active trypsin compared to the warm active counterpart. Electrostatic interactions thus provide an efficient strategy for cold adaptation of trypsin.  相似文献   

16.
In the present study, we investigate the impact of a tightly bound water molecule on ligand binding in the S1 pocket of thrombin. The S1 pocket contains a deeply buried deprotonated aspartate residue (Asp189) that is, due to its charged state, well hydrated in the uncomplexed state. We systematically studied the importance of this water molecule by evaluating a series of ligands that contains pyridine-type P1 side chains that could potentially alter the binding properties of this water molecule. All of the pyridine derivatives retain the original hydration state albeit sometimes with a slight perturbance. In order to prevent a direct H-bond formation with Asp189, and to create a permanent positive charge on the P1 side chain that is positioned adjacent to the Asp189 carboxylate anion, we methylated the pyridine nitrogen. This methylation resulted in displacement of water but was accompanied by a loss in binding affinity. Quantum chemical calculations of the ligand solvation free energy showed that the positively charged methylpyridinium derivatives suffer a large penalty of desolvation upon binding. Consequently, they have a substantially less favorable enthalpy of binding. In addition to the ligand desolvation penalty, the hydration shell around Asp189 has to be overcome, which is achieved in nearly all pyridinium derivatives. Only for the ortho derivative is a partial population of a water next to Asp189 found. Possibly, the gain of electrostatic interactions between the charged P1 side chain and Asp189 helps to compensate for the desolvation penalty. In all uncharged pyridine derivatives, the solvation shell remains next to Asp189, partly mediating interactions between ligand and protein. In the case of the para-pyridine derivative, a strongly disordered cluster of water sites is observed between ligand and Asp189.  相似文献   

17.
The electrostatic potential of Escherichia coli dihydrofolate reductase   总被引:2,自引:0,他引:2  
Escherichia coli dihydrofolate reductase (DHFR) carries a net charge of -10 electrons yet it binds ligands with net charges of -4 (NADPH) and -2 (folate or dihydrofolate). Evaluation and analysis of the electrostatic potential of the enzyme give insight as to how this is accomplished. The results show that the enzyme is covered by an overall negative potential (as expected) except for the ligand binding sites, which are located inside "pockets" of positive potential that enable the enzyme to bind the negatively charged ligands. The electrostatic potential can be related to the asymmetric distribution of charged residues in the enzyme. The asymmetric charge distribution, along with the dielectric boundary that occurs at the solvent-protein interface, is analogous to the situation occurring in superoxide dismutase. Thus DHFR is another case where the shape of the active site focuses electric fields out into solution. The positive electrostatic potential at the entrance of the ligand binding site in E. coli DHFR is shown to be a direct consequence of the presence of three positively charged residues at positions 32, 52, and 57--residues which have also been shown recently to contribute significantly to electronic polarization of the ligand folate. The latter has been postulated to be involved in the catalytic process. A similar structural motif of three positively charged amino acids that gives rise to a positive potential at the entrance to the active site is also found in DHFR from chicken liver, and is suggested to be a common feature in DHFRs from many species. It is noted that, although the net charges of DHFRs from different species vary from +3 to -10, the enzymes are able to bind the same negatively charged ligands, and perform the same catalytic function.  相似文献   

18.
Increase in KC1 concentration from 0.1 m to 3 m enhances the chymotrypsin-catalyzed acyl-transfer to a series of dipeptides, H2N—CH(R1)C(O)NHCH(R2)C(O)O, by factors between 15 and 44. The observed positive salt effect seems to be a screening of the unfavorable electrostatic interaction between the dipeptide carboxyl group and a negatively charged area in the S' subsite on the enzyme surface. The effect is of practical use in preparative peptide synthesis—in 3 m KCl solutions the analytical yields of tripeptides of up to 99 per cent have been obtained.  相似文献   

19.
Short chain fatty acids (SCFAs) are produced in the gut by bacterial fermentation of poorly digested carbohydrates. A key mediator of their actions is the G protein-coupled free fatty acid 2 (FFA2) receptor, and this has been suggested as a therapeutic target for the treatment of both metabolic and inflammatory diseases. However, a lack of understanding of the molecular determinants dictating how ligands bind to this receptor has hindered development. We have developed a novel radiolabeled FFA2 antagonist to probe ligand binding to FFA2, and in combination with mutagenesis and molecular modeling studies, we define how agonist and antagonist ligands interact with the receptor. Although both agonist and antagonist ligands contain negatively charged carboxylates that interact with two key positively charged arginine residues in transmembrane domains V and VII of FFA2, there are clear differences in how these interactions occur. Specifically, although agonists require interaction with both arginine residues to bind the receptor, antagonists require an interaction with only one of the two. Moreover, different chemical series of antagonist interact preferentially with different arginine residues. A homology model capable of rationalizing these observations was developed and provides a tool that will be invaluable for identifying improved FFA2 agonists and antagonists to further define function and therapeutic opportunities of this receptor.  相似文献   

20.
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