首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Enzymes involved in the mammalian microsomal metabolism of drugs are, in numerous cases, inhibited by compounds bearing an imidazolyl scaffold. However, the inhibition potency is highly dependent upon the accessibility of the imidazolyl nitrogen lone pair. In order to highlight some structural parameters of inhibitors that control this phenomenon, a series of compounds containing a nitrogen unsubstituted imidazolyl moiety with varying degrees of nitrogen lone pair accessibility was tested on human and rat hepatic cytochromes P450 and microperoxidase 8, an enzymatically active peptide derived from cytochrome c. In each case, we have shown that the accessibility of the imidazole lone pair determined the extent of inhibition. Nitrogen accessibility was tuned not only by varying the steric hindrance in the vicinity of the imidazolyl ring but also by modifying its surrounding hydrogen bonding network. Compounds in which there exists intramolecular hydrogen bonding between the imidazole moiety and an H-bond acceptor, such as an appropriately positioned amide carbonyl group, demonstrated enhanced inhibitory effects. Conversely, imidazole moieties which are in proximity to H-bond donors, such as an amide NH group, displayed reduced potency. This trend was observed in cyclo-peptide derivatives in which the intramolecular H-bond network was adjusted through the modification of the stereochemistry of a dehydrohistidine residue. It was observed that (Z)-isomers weakly bind heme, whereas (E)-isomers demonstrated higher degrees of metal binding. Therefore, enzymatic inhibition of heme-containing proteins by compounds bearing a dehydrohistidine motif seems to be closely related to its stereochemistry and hydrogen binding propensity. At neutral pH, these differences in binding affinities can be confidently attributed to the ambident H-bond properties of imidazole nitrogen atoms. This structure-activity relationship may be of use for the design of novel imidazolyl compounds as new P450 inhibitors or drug candidates.  相似文献   

2.
The modular synthesis of pyrimidine oligohydrazides and their peptide binding ability are reported. Ethylene glycol substituents ensure water solubility of the compounds. The pattern of hydrogen bond donors and hydrogen bond acceptors resembles the functionalities of a peptide backbone, and intramolecular hydrogen bonds restrict conformational mobility. The pyrimidine heterocycles show emission at 423 nm if either excited with light of 320 nm or by a FRET process from a nearby Trp residue. This property is useful for the luminescent detection of interactions with peptides and proteins.  相似文献   

3.
C K Woodward  L M Ellis 《Biochemistry》1975,14(15):3419-3423
The hydrogen exchange kinetics of the complex of trypsin-soybean trypsin inhibitor (Kunitz) have been compared to the calculated sum of the exchange kinetics for the inhibitor and trypsin measured separately. The exchange rates observed for the complex are substantially less than the sum of the exchange rates in the two individual proteins. These results cannot be accounted for by changes in intermolecular or intramolecular hydrogen bonding. The decrease in exchange rates in the complex are ascribed to changes in solvent accessibility in the component proteins.  相似文献   

4.
A comprehensive structural analysis of X--H...pi hydrogen bonding in proteins is performed based on 592 published high-resolution crystal structures (< or = 1.6 A). All potential donors and acceptors are considered, including acidic C--H groups. The sample contains 1311 putative X--H...pi hydrogen bonds with N--H, O--H or S--H donors, that is about one per 10.8 aromatic residues. By far the most efficient pi-acceptor is the side-chain of Trp, which accepts one X--H...pi hydrogen bond per 5.7 residues. The focus of the analysis is on recurrent structural patterns involving regular secondary structure elements. Numerous examples are found where peptide X--H...pi interactions are functional in stabilization of helix termini, strand ends, strand edges, beta-bulges and regular turns. Side-chain X--H...pi hydrogen bonds are formed in considerable numbers in alpha-helices and beta-sheets. Geometrical data on various types of X--H...pi hydrogen bonds are given.  相似文献   

5.
This paper describes the use of functionalized oligoanthranilamides as conformationally controlled scaffolds for molecular recognition. Oligomers of anthranilamides are stabilized by the formation of intramolecular six-membered hydrogen bonds in a linear strand conformation. Onto alternate anthranilic acid units, we have attached di- or tripeptide recognition units with the potential to form intramolecular hydrogen bonds to an intercalated peptide strand. Using 1H NMR dilution experiments in CDCl(3), we have observed chemical shift changes that are consistent with the formation of an extended hydrogen bonded sheet dimer. We also demonstrate that the bis-alanine functionalized strands are able to form discrete hydrogen bonded complexes with dipeptide substrates and to bind hexanoyl alanylalanine selectively over its benzyl ester. In the presence of excess hydrogen bond donors and acceptors, the oligoanthranilamide strand retained its linear conformation, pointing to the potential of this modular design as a useful and stable scaffold for molecular recognition studies.  相似文献   

6.
The modes of hydrogen bonding of arginine, asparagine, and glutamine side chains and of urea have been examined in small-molecule crystal structures in the Cambridge Structural Database and in crystal structures of protein-nucleic acid and protein-protein complexes. Analysis of the hydrogen bonding patterns of each by graph-set theory shows three patterns of rings (R) with one or two hydrogen bond acceptors and two donors and with eight, nine, or six atoms in the ring, designated R2 2(8), R2 2(9), and R1 2(6). These three patterns are found for arginine-like groups and for urea, whereas only the first two patterns Rl(8) and R2(9) are found for asparagine- and glutamine-like groups. In each case, the entire system is planar within 0.7 Å or less. On the other hand, in macromolecular crystal structures, the hydrogen bonding patterns in protein-nucleic acid complexes between the nucleic acid base and the protein are all R2 2(9), whereas hydrogen bonding between Watson-Crick-like pairs of nucleic acid bases is R2 2(8). These two hydrogen bonding arrangements [R2 2(9) and R2 2(8)] are predetermined by the nature of the groups available for hydrogen bonding. The third motif identified, R1 2(6), involves hydrogen bonds that are less linear than in the other two motifs and is found in proteins.  相似文献   

7.
We have developed a free‐energy function based on an all‐atom model for proteins. It comprises two components, the hydration entropy (HE) and the total dehydration penalty (TDP). Upon a transition to a more compact structure, the number of accessible configurations arising from the translational displacement of water molecules in the system increases, leading to a water‐entropy gain. To fully account for this effect, the HE is calculated using a statistical‐mechanical theory applied to a molecular model for water. The TDP corresponds to the sum of the hydration energy and the protein intramolecular energy when a fully extended structure, which possesses the maximum number of hydrogen bonds with water molecules and no intramolecular hydrogen bonds, is chosen as the standard one. When a donor and an acceptor (e.g., N and O, respectively) are buried in the interior after the break of hydrogen bonds with water molecules, if they form an intramolecular hydrogen bond, no penalty is imposed. When a donor or an acceptor is buried with no intramolecular hydrogen bond formed, an energetic penalty is imposed. We examine all the donors and acceptors for backbone‐backbone, backbone‐side chain, and side chain‐side chain intramolecular hydrogen bonds and calculate the TDP. Our free‐energy function has been tested for three different decoy sets. It is better than any other physics‐based or knowledge‐based potential function in terms of the accuracy in discriminating the native fold from misfolded decoys and the achievement of high Z‐scores. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

8.
Hydrogen bonding is a key contributor to the specificity of intramolecular and intermolecular interactions in biological systems. Here, we develop an orientation-dependent hydrogen bonding potential based on the geometric characteristics of hydrogen bonds in high-resolution protein crystal structures, and evaluate it using four tests related to the prediction and design of protein structures and protein-protein complexes. The new potential is superior to the widely used Coulomb model of hydrogen bonding in prediction of the sequences of proteins and protein-protein interfaces from their structures, and improves discrimination of correctly docked protein-protein complexes from large sets of alternative structures.  相似文献   

9.
For 30 years, the prevailing view has been that the hydrophobic effect contributes considerably more than hydrogen bonding to the conformational stability of globular proteins. The results and reasoning presented here suggest that hydrogen bonding and the hydrophobic effect make comparable contributions to the conformational stability of ribonuclease T1 (RNase T1). When RNase T1 folds, 86 intramolecular hydrogen bonds with an average length of 2.95 A are formed. Twelve mutants of RNase T1 [Tyr----Phe (5), Ser----Ala (3), and Asn----Ala (4)] have been prepared that remove 17 of the hydrogen bonds with an average length of 2.93 A. On the basis of urea and thermal unfolding studies of these mutants, the average decrease in conformational stability due to hydrogen bonding is 1.3 kcal/mol per hydrogen bond. This estimate is in good agreement with results from several related systems. Thus, we estimate that hydrogen bonding contributes about 110 kcal/mol to the conformational stability of RNase T1 and that this is comparable to the contribution of the hydrophobic effect. Accepting the idea that intramolecular hydrogen bonds contribute 1.3 +/- 0.6 kcal/mol to the stability of systems in an aqueous environment makes it easier to understand the stability of the "molten globule" states of proteins, and the alpha-helical conformations of small peptides.  相似文献   

10.
Hydrogen-bonding effects and 13C-NMR of the DNA double helix   总被引:1,自引:1,他引:0       下载免费PDF全文
13C-nmr chemical shifts of the nucleotides in DNA are sensitive to hydrogen bonding, especially for three of the carbons immediately bonded to exocyclic oxygen or nitrogen atoms acting as H-bond acceptors or donors. GuoC2, GuoC6 and ThdC4 are strongly deshielded (about 1 ppm) upon Watson-Crick pairing in oligodeoxynucleotide duplexes, regardless of the base sequence. Deshielding at these sites may be useful to distinguish bases involved in Watson-Crick pairs from unpaired bases.  相似文献   

11.
Non-traditional C-H cdots, three dots, centered Y hydrogen bonds, in which a carbon atom acts as the hydrogen donor and an electronegative atom Y (Y=N, O or S) acts as the acceptor, have been reported in proteins, but their importance in protein structures is not well established. Here, we present the results of three computational tests that examine the significance of C-H cdots, three dots, centered Y bonds involving the C(alpha) in proteins. First, we compared the number of C(alpha)-H cdots, three dots, centered Y bonds in native structures with two sets of compact, energy-minimized decoy structures. The decoy structures contain about as many C(alpha)-H cdots, three dots, centered Y bonds as the native structures, indicating that the constraints of chain connectivity and compactness can lead to incidental formation of C(alpha)-H cdots, three dots, centered Y bonds. Secondly, we examined whether short C(alpha)-H cdots, three dots, centered Y bonds have a tendency to be linear, as is expected for a cohesive hydrogen-bonding interaction. The native structures do show this trend, but so does one of the decoy sets, suggesting that this criterion is also not sufficient to indicate a stabilizing interaction. Finally, we examined the preference for C(alpha)-H cdots, three dots, centered Y bond donors to be near to strong hydrogen bond acceptors. In the native proteins, the alpha protons attract strong acceptors like oxygen atoms more than weak acceptors. In contrast, hydrogen bond donors in the decoy structures do not distinguish between strong and weak acceptors. Thus, any individual C(alpha)-H cdots, three dots, centered Y bond may be fortuitous and occur due to the polypeptide connectivity and compactness. Taken collectively, however, C(alpha)-H cdots, three dots, centered Y bonds provide a weakly cohesive force that stabilizes proteins.  相似文献   

12.
Summary Temperature coefficients are widely used as an indication of solvent accessibility to amide protons. Low temperature coefficients are related to low accessibility and are often interpreted as evidence for intramolecular hydrogen bonding. Conformational shifts, i.e. the difference between chemical shifts of a particular residue in a structured and in a random-coil conformation, provide information on secondary structure. In particular, negative CHα conformational shifts are often used to delineate the extent of helical stretches. NH conformational shifts show large oscillations within a helix that have been interpreted as the result of helix distortions affecting hydrogen bond lengths. In the ocurse of the study of differnet peptides that adopt a helical structure in the presence of the structure-inducing solvent hexafluoroisopropanol (HFIP), we have found a strong correlation between temperature coefficients and amide conformational shifts. However, contrary to the initial expectations, lower temperature coefficients were associated to amide protons involved in longer, and presumably weaker, hydrogen bonds. The correlation can be explained, however, assuming that, in helical peptides dissolved in HFIP, temperature affects the chemical shift of amide protons mainly by changing the average length of intramolecular hydrogen bonds and changes in solvent accessibility play only a secondary role under these experimental conditions. The pattern of temperature coefficients in helical peptides can therefore be used to identify short or long hydragen bonds causing bending of the helix axis.  相似文献   

13.
Membrane-bound proteins owe their retention and conformation in the lipid bilayer to hydrophobic peptide domains. Additional fixation, by protein-lipid hydrogen bonding, has been suggested, and recent reports on protein kinase C activation by diacylglycerol (DG) provide an unambiguous model for such bonding. The sn-1,2-diacylglycerol appears to donate a hydrogen bond from the sn-3 hydroxyl to the enzyme and to receive two hydrogen bonds, in the sn-1 and sn-2 ester CO groups, from the enzyme. This arrangement is confirmed in phorbol ester, a competitive inhibitor of DG for the kinase. This tumor promotor has a nearly identical spatial arrangement of hydrogen bond donor (9 alpha-OH) and acceptors (12 and 13 ester CO); so have two other tumor promotors, teleocidin and aplysiatoxin. There are reasons to believe that protein kinase C is not the only protein that is bound to membrane lipids by hydrogen bonding, and such bonding will have to be considered in membrane-associated events such as fusion, cross-membrane transport, or anesthesia.  相似文献   

14.
《Inorganica chimica acta》1986,123(3):155-159
The crystal structure analysis of the 2:1 adduct of cyclosarcosylsarcosine with copper(II) perchlorate shows that the independent unit is composed of six water molecules octahedrally coordinated to the Cu(II) ion, two tetrahedral perchlorate ions and four independent halves of cyclosarcosylsarcosine molecules lying on crystallographic centers of symmetry. All available hydrogens of water molecules are involved in hydrogen bonding as donors and all carbonyl oxygen atoms of the cyclic peptide molecules function as acceptors. Two other oxygen atoms for each perchlorate anion participate in the hydrogen bonding scheme, which leads to the absence of the orientational disorder usually observed in these ions. Layers of inorganic material and layers of organic material, roughly parallel to the ab plane, pack alternatively with each other. Electrostatic and ion-dipole interactions together with hydrogen bonds are responsible for the building up of the crystals.  相似文献   

15.
Temperature coefficients are widely used as an indication of solvent accessibility to amide protons. Low temperature coefficients are related to low accessibility and are often interpreted as evidence for intramolecular hydrogen bonding. Conformational shifts, i.e. the difference between chemical shifts of a particular residue in a structured and in a random-coil conformation, provide information on secondary structure. In particular, negative CH conformational shifts are often used to delineate the extent of helical stretches. NH conformational shifts show large oscillations within a helix that have been interpreted as the result of helix distortions affecting hydrogen bond lengths. In the course of the study of different peptides that adopt a helical structure in the presence of the structure-inducing solvent hexafluoroisopropanol (HFIP), we have found a strong correlation between temperature coefficients and amide conformational shifts. However, contrary to the initial expectations, lower temperature coefficients were associated to amide protons involved in longer, and presumably weaker, hydrogen bonds. The correlation can be explained, however, assuming that, in helical peptides dissolved in HFIP, temperature affects the chemical shift of amide protons mainly by changing the average length of intramolecular hydrogen bonds and changes in solvent accessibility play only a secondary role under these experimental conditions. The pattern of temperature coefficients in helical peptides can therefore be used to identify short or long hydrogen bonds causing bending of the helix axis.  相似文献   

16.
A series of arylidene anabaseines were synthesized to probe the functional impact of hydrogen bonding on human α7 nicotinic acetylcholine receptor (nAChR) activation and desensitization. The aryl groups were either hydrogen bond acceptors (furans), donors (pyrroles), or neither (thiophenes). These compounds were tested against a series of point mutants of the ligand-binding domain residue Gln-57, a residue hypothesized to be proximate to the aryl group of the bound agonist and a putative hydrogen bonding partner. Q57K, Q57D, Q57E, and Q57L were chosen to remove the dual hydrogen bonding donor/acceptor ability of Gln-57 and replace it with hydrogen bond donating, hydrogen bond accepting, or nonhydrogen bonding ability. Activation of the receptor was compromised with hydrogen bonding mismatches, for example, pairing a pyrrole with Q57K or Q57L, or a furan anabaseine with Q57D or Q57E. Ligand co-applications with the positive allosteric modulator PNU-120596 produced significantly enhanced currents whose degree of enhancement was greater for 2-furans or -pyrroles than for their 3-substituted isomers, whereas the nonhydrogen bonding thiophenes failed to show this correlation. Interestingly, the PNU-120596 agonist co-application data revealed that for wild-type α7 nAChR, the 3-furan desensitized state was relatively stabilized compared with that of 2-furan, a reversal of the relationship observed with respect to the barrier for entry into the desensitized state. These data highlight the importance of hydrogen bonding on the receptor-ligand state, and suggest that it may be possible to fine-tune features of agonists that mediate state selection in the nAChR.  相似文献   

17.
Hydrogen bonding stabilizes globular proteins.   总被引:8,自引:1,他引:7       下载免费PDF全文
It is clear that intramolecular hydrogen bonds are essential to the structure and stability of globular proteins. It is not clear, however, whether they make a net favorable contribution to this stability. Experimental and theoretical studies are at odds over this important question. Measurements of the change in conformational stability, delta (delta G), for the mutation of a hydrogen bonded residue to one incapable of hydrogen bonding suggest a stabilization of 1.0 kcal/mol per hydrogen bond. If the delta (delta G) values are corrected for differences in side-chain hydrophobicity and conformational entropy, then the estimated stabilization becomes 2.2 kcal/mol per hydrogen bond. These and other experimental studies discussed here are consistent and compelling: hydrogen bonding stabilizes globular proteins.  相似文献   

18.
The molecular recognition and discrimination of adenine and guanine ligand moieties in complexes with proteins have been studied using empirical observations on carefully selected crystal structures. The distribution of protein folds that bind these purines has been found to differ significantly from that across the whole PDB, but the most populated architectures and folds are also the most common in three genomes from the three different domains of life. The protein environments around the two nucleic acid bases were significantly different, in terms of the propensities of amino acid residues to be in the binding site, as well as their propensities to form hydrogen bonds to the bases. Plots of the distribution of protein atoms around the two purines clearly show different clustering of hydrogen bond donors and acceptors opposite complimentary acceptors and donors in the rings, with hydrophobic areas below and above the rings. However, the clustering pattern is fuzzy, reflecting the variety of ways that proteins have evolved to recognise the same molecular moiety. Furthermore, an analysis of the conservation of residues in the protein chains binding guanine shows that residues in contact with the base are in general better conserved than the rest of the chain.  相似文献   

19.
Recognition of a DNA sequence by a protein is achieved by interface-coupled chemical and shape complementation. This complementation between the two molecules is clearly directional and is determined by the specific chemical contacts including mainly hydrogen bonds. Directionality is an instrumental property of hydrogen bonding as it influences molecular conformations, which also affects DNA-protein recognition. The prominent elements in the recognition of a particular DNA sequence by a protein are the hydrogen-bond donors and acceptors of the base pairs into the grooves of the DNA that must interact with complementary moieties of the protein partner. Protein side chains make most of the crucial contacts through bidentate and complex hydrogen-bonding interactions with DNA base edges hence conferring remarkable specificity.  相似文献   

20.
Brieba LG  Sousa R 《Biochemistry》2000,39(5):919-923
On the basis of their recently described T7 RNA polymerase-T7 promoter crystal structure, Cheetham et al. [(1999) Nature 399, 80] propose that discrimination of the hydrogen bonding character of the elongating NTP ribose 2'-substituent involves a hydrogen bond to histidine 784. This would contradict a previous conclusion, based on the characterization of mutant RNAPs, that discrimination of the hydrogen bonding character of the ribose 2'-substituent depends solely on the hydroxyl group of tyrosine 639. To resolve this point, we prepared and characterized histidine 784 point mutants. We find that while these mutations reduce the activity of the polymerase, they do not significantly reduce the level of ribose discrimination. Furthermore, a mutant with alanine at position 784 preferentially utilizes NTPs with 2'-substituents capable of acting as hydrogen bond donors or acceptors (2'-OH and 2'-NH(2)) over NTPs with substituents that lack such properties (2'-F and 2'-H). In contrast, mutation of tyrosine 639 to phenylalanine eliminates discrimination of ribose 2'-group hydrogen bonding character. The effects on ribose discrimination of mutating tyrosine 639 to phenylalanine are independent of the side chain at position 784. These results indicate that histidine 784 is not involved in discrimination of the ribose 2'-group of the elongating NTP. The ability of T7RNAP tyrosine 639, which is conserved in both RNA and DNA polymerases, to select for rNTPs appears to be due to the fact that in RNAPs this tyrosine is available to hydrogen bond to the ribose 2'-OH, while in DNAPs it is hydrogen bonded to a glutamic acid.  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