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 0 毫秒
1.
Recent site‐resolved hydrogen exchange measurements have uncovered significant discrepancies between simulations and experimental data during protein folding, including the excessive intramolecular hydrogen bonds in simulations. This finding indicates a possibility that intramolecular charge–charge interactions have not included sufficient dielectric screening effect of the electronic polarization. Scaling down peptide atomic charges according to the optical dielectric constant is tested in this study. As a result, the number of intramolecular hydrogen bonds is lower than using unscaled atomic charges while reaching the same levels of helical contents or β‐hairpin backbone hydrogen bonds, because van der Waals interactions contribute substantially to peptide folding in water. Reducing intramolecular charge–charge interactions and hydrogen bonding increases conformational search efficiency. In particular, it reduces the equilibrium helical content in simulations using AMBER force field and the energy barrier in folding simulations using CHARMM force field.  相似文献   

2.
We present an ab initio molecular dynamics study of the roles of fluctuating hydrogen bonds and free ND modes in the dynamics of ND stretch frequency fluctuations in deuterated liquid ammonia. We have also looked at some of the other dynamical quantities such as diffusion and orientational relaxation and also structural quantities such as pair correlations and hydrogen bonding properties which are relevant in the current context. The time correlation function of ND stretch frequencies is found to decay with primarily two time scales: A short-time decay with a time scale of less than 100 fs arising from intermolecular motion of intact hydrogen bonds and also from fast hydrogen bond breaking and a longer time scale of about 500 fs which can be assigned to the lifetime of free ND modes. Unlike water, in liquid ammonia an ND mode is found to remain free for a longer period than it stays hydrogen bonded and this longer lifetime of free ND modes determines the long-time behaviour of frequency fluctuations. Our hole dynamics calculations produced results of vibrational spectral diffusion that are similar to the decay of frequency time correlation. Inclusion of dispersion corrections is found to make the dynamics slightly faster.  相似文献   

3.
Our goal was to gain a better understanding of the contribution of the burial of polar groups and their hydrogen bonds to the conformational stability of proteins. We measured the change in stability, Δ(ΔG), for a series of hydrogen bonding mutants in four proteins: villin headpiece subdomain (VHP) containing 36 residues, a surface protein from Borrelia burgdorferi (VlsE) containing 341 residues, and two proteins previously studied in our laboratory, ribonucleases Sa (RNase Sa) and T1 (RNase T1). Crystal structures were determined for three of the hydrogen bonding mutants of RNase Sa: S24A, Y51F, and T95A. The structures are very similar to wild type RNase Sa and the hydrogen bonding partners form intermolecular hydrogen bonds to water in all three mutants. We compare our results with previous studies of similar mutants in other proteins and reach the following conclusions. (1) Hydrogen bonds contribute favorably to protein stability. (2) The contribution of hydrogen bonds to protein stability is strongly context dependent. (3) Hydrogen bonds by side chains and peptide groups make similar contributions to protein stability. (4) Polar group burial can make a favorable contribution to protein stability even if the polar groups are not hydrogen bonded. (5) The contribution of hydrogen bonds to protein stability is similar for VHP, a small protein, and VlsE, a large protein.  相似文献   

4.
It is often assumed that the peptide backbone forms a substantial number of additional hydrogen bonds when a protein unfolds. We challenge that assumption in this article. Early surveys of hydrogen bonding in proteins of known structure typically found that most, but not all, backbone polar groups are satisfied, either by intramolecular partners or by water. When the protein is folded, these groups form approximately two hydrogen bonds per peptide unit, one donor or acceptor for each carbonyl oxygen or amide hydrogen, respectively. But when unfolded, the backbone chain is often believed to form three hydrogen bonds per peptide unit, one partner for each oxygen lone pair or amide hydrogen. This assumption is based on the properties of small model compounds, like N-methylacetamide, or simply accepted as self-evident fact. If valid, a chain of N residues would have approximately 2N backbone hydrogen bonds when folded but 3N backbone hydrogen bonds when unfolded, a sufficient difference to overshadow any uncertainties involved in calculating these per-residue averages. Here, we use exhaustive conformational sampling to monitor the number of H-bonds in a statistically adequate population of blocked polyalanyl-six-mers as the solvent quality ranges from good to poor. Solvent quality is represented by a scalar parameter used to Boltzmann-weight the population energy. Recent experimental studies show that a repeating (Gly-Ser) polypeptide undergoes a denaturant-induced expansion accompanied by breaking intramolecular peptide H-bonds. Results from our simulations augment this experimental finding by showing that the number of H-bonds is approximately conserved during such expansion⇋compaction transitions.  相似文献   

5.
Hydrogen bonding pattern of nucleotides and carbohydrates has been analysed using Cambridge database. An analysis on ribonucleotides shows the 3′ …5′ hydrogen bond mediated aggregation to be the most common alignment. The 2′ …5′ alignment, which occurs under special circumstances in nature, is found to be the second choice. An analysis on carbohydrates suggests that self assembly of these molecules is not conducive to the formation of polysaccharides of the type which are found in present day living organisms. Further, the role of water molecules in the polymerization of three important biomolecules, namely nucleotides, carbohydrates and amino acids, has been analysed. Implication of these results to prebiotic polymerization is discussed DCB contribution No. 804.  相似文献   

6.
It was shown earlier, from database analysis, model building studies, and molecular dynamics simulations that formation of cross-strand bifurcated or Extra Watson-Crick hydrogen (EWC) bonds between successive base pairs may lead to extra rigidity to DNA double helices of certain sequences. The strengths of these hydrogen bonds are debatable, however, as they do not have standard linear geometry criterion. We have therefore carried out detailed ab initio quantum chemical studies using RHF/6-31G(2d,2p) and B3LYP/6-31G(2p,2d) basis sets to determine strengths of several bent hydrogen bonds with different donor and acceptors. Interaction energy calculations, corrected for the basis set superposition errors, suggest that N-H...O type bent EWC hydrogen bonds are possible along same strands or across the strands between successive base pairs, leading to significant stability (ca. 4-9 kcal/mol). The N-H...N and C-H...O type interactions, however, are not so stabilizing. Hence, consideration of EWC N-H...O H-bonds can lead to a better understanding of DNA sequence directed structural features.  相似文献   

7.
Analysis of 100 complete sets of the cytoplasmic elongator tRNA genes from Bacteria, Archaea, and Eukarya pointed to correspondences between types of anticodon and composition of the rest of the tRNA body. The number of the hydrogen bonds formed between the complementary nucleotides in the anticodon-codon duplex appeared as a major quantitative parameter determining covariations in all three domains of life. Our analysis has supported and advanced the "extended anticodon" concept that is based on the argument that the decoding performance of the anticodon is enhanced by selection of a matching anticodon stem-loop sequence, as reported by Yarus in 1982. In addition to the anticodon stem-loop, we have found covariations between the anticodon nucleotides and the composition of the distant regions of their respective tRNAs that include dihydrouridine (D) and thymidyl (T) stem-loops. The majority of the covariable tRNA positions were found at the regions with the increased dynamic potential--such as stem-loop and stem-stem junctions. The consistent occurrences of the covariations on the multigenomic level suggest that the number and pattern of the hydrogen bonds in the anticodon-codon duplex constitute a major factor in the course of translation that is reflected in the fine-tuning of the tRNA composition and structure.  相似文献   

8.
Peters D  Peters J 《Biopolymers》2004,73(2):178-191
A study of the role of the hydrogen-bonding side chains in the ribbon of hydrogen bonds in globular proteins, using the papain family as an example, suggests that these side chains may be divided into three categories depending on their position in the molecule. In the first category, they form part of the local ribbon, in the second they form part of the ribbon at a site remote along the main chain, and in the third they play no role in the formation of the ribbon. The second case is particularly interesting because it provides a natural mechanism for the formation of the tertiary structure of the globular proteins. The results suggest that the robustness of the globular proteins towards mutations arises from the fact that many mutations that involve hydrogen-bonding side chains either leave the hydrogen bonding of the ribbon essentially unchanged or their hydrogen bonding plays no part in the formation of the ribbon in the first place. The results show that it is possible to obtain the ribbon of hydrogen bonds for a family of proteins whose data set's are of intermediate quality by studying the ribbons of several members of such a family and then taking an average over the different partial ribbons to create a standard ribbon of hydrogen bonds for the family as a whole. This method is used here to derive the standard ribbon for the papain family with papain itself, actinidin, and human liver cathepsin B as the representatives of the family. All three members of the family fit the standard ribbon with an accuracy of 85-91%. This result opens up the use of this technique for the study of a large number of globular proteins whose recorded data sets are of intermediate quality.  相似文献   

9.
Petrella RJ  Karplus M 《Proteins》2004,54(4):716-726
Although most side-chain torsion angles correspond to low-energy rotameric positions, deviations occur with significant frequency. One striking example arises in Trp residues, which have an important role in stabilizing protein structures because of their size and mixed hydrophobic/hydrophilic character. Ten percent of Trp side-chains have unexplained conformations with chi(2) near 0 degrees instead of the expected 90 degrees. The current work is a structural and energetic analysis of these conformations. It is shown that many Trp residues with these orientations are stabilized by three-center carbon-donor hydrogen bonds of the form C-H...X...H-C, where X is a polar hydrogen-bond acceptor in the environment of the side-chain. The bridging hydrogen bonds occur both within the Trp side-chain and between the side-chain and the local protein backbone. Free energy maps of an isolated Trp residue in an explicit water environment show a minimum corresponding to the off-rotamer peak observed in the crystallographic data. Bridging carbon-donor hydrogen bonds are also shown to stabilize on-rotamer Trp conformations, and similar bridging hydrogen bonds also stabilize some off-rotamer Asp conformations. The present results suggest a previously unrecognized role for three-center carbon-donor hydrogen bonds in protein structures and support the view that the off-rotamer Trp side-chain orientations are real rather than artifacts of crystallographic refinements. Certain of the off-rotamer Trp conformations appear to have a functional role.  相似文献   

10.
Yuanyuan Li 《Luminescence》2017,32(8):1385-1390
A ratiometric fluorescent chemosensor 2‐(2′‐hydroxy‐3′‐formyl‐5′‐methoxyphenyl)benzothiazole ( 1 ) was developed for the detection of cysteine (Cys). In aqueous solution at neutral pH, 1 exhibited a ratiometric fluorescent response to Cys with a remarkable red‐to‐green shift in the emission wavelength. This fluorescence change was attributed to the cyclization reaction between the formyl group in 1 and the amino and sulfhydryl group in Cys in a stoichiometry of 1: 1 according to the proposed mechanism. At neutral pH, 1 displayed a significant fluorescence ratio signal enhancement with the addition of Cys. Furthermore, 1 showed good selectivity toward Cys. The detection limit and linear range were 5.6 and 0–100 μmol/L, respectively, which demonstrated that 1 could recognize relatively low concentrations of Cys and is a good candidate for applications in detecting Cys.  相似文献   

11.
We prepared an aminothiourea‐derived Schiff base (DA) as a fluorescent chemosensor for Hg2+ ions. Addition of 1 equiv of Hg2+ ions to the aqueous solution of DA gave rise to an obvious fluorescence enhancement and the subsequent addition of more Hg2+ induced gradual fluorescence quenching. Other competing ions, including Pb2+, Cd2+, Cr3+, Zn2+, Fe2+, Co3+, Ni2+, Ca2+, Mg2+, K+ and Na+, did not induce any distinct fluorescence changes, indicating that DA can selectively detect Hg2+ ions in aqueous solution. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

12.
Hydrogen bonding networks proximal to metal centers are emerging as a viable means for controlling secondary coordination spheres. This has led to the regulation of reactivity and isolation of complexes with new structural motifs. We have used the tridenate ligand bis[(N′-tert-butylureido)-N-ethyl]-N-methylaminato ([H21]2−) that contains two hydrogen bond donors to examine the oxidation of the FeII-acetate complex, [FeIIH212-OAc)] with dioxygen, amine N-oxides, and xylyl azide. A complex with FeIII-O-FeIII core results from the oxidation with dioxygen and amine N-oxides, in which the oxo ligand is involved in hydrogen bonding to the [H21]2− ligand. A distinctly different hydrogen bonding network was found in FeIII dimer isolated from the reaction with the xylyl azide: a rare FeIII-N(R)-FeIII core was observed that does not have hydrogen bonds to the bridging nitrogen atom. The intramolecular H-bond networks within these dimers appear to adjust to the presence of the bridging species and rearrange to its size and electron density.  相似文献   

13.
14.
Adamian L  Liang J 《Proteins》2002,47(2):209-218
Polar and ionizable amino acid residues are frequently found in the transmembrane (TM) regions of membrane proteins. In this study, we show that they help to form extensive hydrogen bond connections between TM helices. We find that almost all TM helices have interhelical hydrogen bonding. In addition, we find that a pair of contacting TM helices is packed tighter when there are interhelical hydrogen bonds between them. We further describe several spatial motifs in the TM regions, including "Polar Clamp" and "Serine Zipper," where conserved Ser residues coincide with tightly packed locations in the TM region. With the examples of halorhodopsin, calcium-transporting ATPase, and bovine cytochrome c oxidase, we discuss the roles of hydrogen bonds in stabilizing helical bundles in polytopic membrane proteins and in protein functions.  相似文献   

15.
Xia J  Case DA 《Biopolymers》2012,97(5):276-288
Although the crystal structure of the disaccharide sucrose was solved more than 30 years ago, its conformational distribution in aqueous solution is still a matter of debate. We report here a variety of molecular dynamics simulations (mostly of 100 ns) using the GLYCAM06 force field and various water models, paying particular attention to comparisons to NMR measurements of residual dipolar couplings and electron-mediated spin-spin couplings. We focus on the glycosidic linkage conformation, the puckering phase angle of the fructose ring, and intramolecular hydrogen bonds between the two sugars. Our results show that sucrose is indeed a dynamic molecule, but the crystal conformation is qualitatively the dominant one in dilute solution. A second conformational basin, populated in many force fields, is probably overstabilized in the calculations.  相似文献   

16.
NaOH/urea aqueous solution has been used as a solvent for chitin for the first time. Effects of this solvent composition and temperature on the solubility and stability of chitin solution were studied with an optical microscope, from which 8 wt% NaOH/4 wt% urea concentrations were deduced as suitable and −20 °C as the appropriate temperature. The original and regenerated chitin were characterized by viscosity, elemental analysis, FI-IR and X-RD analysis, and the effect of solvent composition and temperature on chitin structure was investigated. It was inferred that 8 wt% NaOH/4 wt% urea solvent under low temperature adventitiously has little effect on chitin structure and the urea is of benefit to the stability of chitin solution. In addition, the rheological properties suggested that chitin aqueous solution in high concentration is a pseudoplastic fluid and that chitin aqueous solution in low concentrations is a Newtonian fluid. This chitin aqueous solution is sensitive to temperature and will transform it to a gel when temperature increases.  相似文献   

17.
In this paper, we demonstrate that the sensitivity of triple-resonance NMR experiments can be enhanced significantly through quenching scalar coupling mediated relaxation by using composite-pulse decoupling (CPD) or an adiabatic decoupling sequence on aliphatic, in particular alpha-carbons in 13C/15N-labeled proteins. The CPD-HNCO experiment renders 50% sensitivity enhancement over the conventional CT-HNCO experiment performed on a 12 kDa FK506 binding protein, when a total of 266 ms of amide nitrogen–carbonyl carbon defocusing and refocusing periods is employed. This is a typical time period for the direct detection of hydrogen bonds in proteins via trans-hydrogen bond 3h J NC couplings. The experimental data fit theoretical analysis well. The significant enhancement in sensitivity makes the experiment more applicable to larger-sized proteins without resorting to perdeuteration.  相似文献   

18.
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   

19.
The peptide toxin bombolitin III [B(III)], originally isolated from bumblebee venom, has been shown to undergo a concentration-dependent conformational change from a random structure to an α-helix induced by aggregation. The aggregation process and the consequent folding results from a delicate balance of electrostatic and hydrophobic interactions. The conformational change is strongly dependent on pH and salt concentration. In order to gain insight on the structure of the aggregates, and in particular, on the aggregation number and relative orientation of helices in the molecular complexes, the following analogue of bombolitin III was designed and synthesized: Ile-Lys-Bpa-Met-Asp-Ile-Leu-Ala-Lys-Leu-Gly-Lys-Val-Leu-Ala-His-Val-NH2 Bpa3-B(III) where Bpa is benzoylphenylalanine. Bpa3-B(III) aggregates were investigated by CD and nmr techniques. The observed nuclear Overhauser effect pattern accounts for an antiparallel orientation of two distinct helices. The Bpa side chain allows for the photoinduced cross reaction with any aliphatic proton in spatial proximity. After irradiation, the reaction mixture was analyzed by high performance liquid chromatography and electrospray mass spectrometry. The results confirmed the presence of dimeric and trimeric complexes of bombolitin III formed upon interhelix cross-linking. © 1997 John Wiley & Sons, Inc. Biopoly 42: 147–156, 1997  相似文献   

20.
Electron microscopy shows that hyaluronan (HA) forms sheets and tube-like structures in solution. Molecular modelling by Tartu plastic space-filling atomic models revealed that hydroxymethyl and carboxylate groups of HA anti-parallel chains can be joined by H-bonds. Using these bonds, HA molecules can be modelled as sheets and tubules. These tertiary structures have three kinds of lateral contact: (1) antiparallel chains stacked by hydrophobic patches; (2) parallel chains joined by both stacking interactions and H-bonds; and (3) crossing chains joined by H-bonds and stacking interactions. Sheet and tubular structures may explain some viscoelastic and biological properties of HA.  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