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 15 毫秒
1.
Molecular Dynamics simulations have been carried out for 100 ps on crystal structure of beta-cyclodextrin in vacuo and with explicit inclusion of solvent at constant pressure and constant temperature using the GROMOS MD algorithm, with a time step of 0.005 ps. The conformational space of the glycosidic linkage was studied by calculating two virtual dihedrals connecting the successive glucose units for the 2000 structures saved during the two simulations. Three preferred regions for alpha 1-4 glycosidic linkage were found in both the simulations. The use of these virtual dihedral angles in representing the glycosidic linkage is also brought out from these studies.  相似文献   

2.
The macromolecular conformation of the constituent polysaccharides in lignocellulosic biomass influences their supramolecular interactions, and therefore their function in plants and their performance in technical products. The flexibility of glycosidic linkages from the backbone of hemicelluloses was studied by evaluating the conformational freedom of the φ and ψ dihedral angles using molecular dynamic simulations, additionally selected molecules were correlated with experimental data by nuclear magnetic resonance spectroscopy. Three types of β‐(1→4) glycosidic linkages involving the monosaccharides (Glcp, Xylp and Manp) present in the backbone of hemicelluloses were defined. Different di‐ and tetrasaccharides with combinations of such sugar monomers from hemicelluloses were simulated, and free energy maps of the φ – ψ space and hydrogen‐bonding patterns were obtained. The glycosidic linkage between Glc‐Glc or Glc‐Man (C‐type) was the stiffest with mainly one probable conformation; the linkage from Man‐Man or Man‐Glc (M‐type) was similar but with an increased probability for an alternative conformation making it more flexible, and the linkage between two Xyl‐units (X‐type) was the most flexible with two almost equally populated conformations. Glycosidic linkages of the same type showed essentially the same conformational space in both disaccharides and in the central region of tetrasaccharides. Different probabilities of glycosidic linkage conformations in the backbone of hemicelluloses can be directly estimated from the free energy maps, which to a large degree affect the overall macromolecular conformations of these polymers. The information gained contributes to an increased understanding of the function of hemicelluloses both in the cell wall and in technical products.  相似文献   

3.
The Ramachandran map clearly delineates the regions of accessible conformational (φ–ψ) space for amino acid residues in proteins. Experimental distributions of φ, ψ values in high‐resolution protein structures, reveal sparsely populated zones within fully allowed regions and distinct clusters in apparently disallowed regions. Conformational space has been divided into 14 distinct bins. Residues adopting these relatively rare conformations are presented and amino acid propensities for these regions are estimated. Inspection of specific examples in a completely “arid”, fully allowed region in the top left quadrant establishes that side‐chain and backbone interactions may provide the energetic compensation necessary for populating this region of φ–ψ space. Asn, Asp, and His residues showed the highest propensities in this region. The two distinct clusters in the bottom right quadrant which are formally disallowed on strict steric considerations correspond to the gamma turn (C7 axial) conformation (Bin 12 ) and the i + 1 position of Type II′ β turns (Bin 13) . Of the 516 non‐Gly residues in Bin 13 , 384 occupied the i + 1 position of Type II′ β turns. Further examination of these turn segments revealed a high propensity to occur at the N‐terminus of helices and as a tight turn in β hairpins. The β strand–helix motif with the Type II′ β turn as a connecting element was also found in as many as 57 examples. Proteins 2014; 82:1101–1112. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
Xia J  Case DA 《Biopolymers》2012,97(5):289-302
We report 100 ns molecular dynamics simulations, at various temperatures, of sucrose in water (with concentrations of sucrose ranging from 0.02 to 4M), and in a 7:3 water‐DMSO mixture. Convergence of the resulting conformational ensembles was checked using adaptive‐biased simulations along the glycosidic Φ and ψ torsion angles. NMR relaxation parameters, including longitudinal (R1) and transverse (R2) relaxation rates, nuclear Overhauser enhancements (NOE), and generalized order parameter (S2) were computed from the resulting time‐correlation functions. The amplitude and time scales of molecular motions change with temperature and concentration in ways that track closely with experimental results, and are consistent with a model in which sucrose conformational fluctuations are limited (with 80–90% of the conformations having ??ψ values within 20° of an average conformation), but with some important differences in conformation between pure water and DMSO‐water mixtures. © 2011 Wiley Periodicals, Inc. Biopolymers 97: 289–302, 2012.  相似文献   

5.
Molecular mechanics and dynamics calculations were carried out on the disaccharides α-L-Rhap-(1 → 2)-α-L-Rhap-(1 → OMe) (1) and α-L-Rhap-(1 → 3)-α-L-Rhap-(1 OMe) (2), and the trisaccharide α-L-Rhap-(1 → 2)-α-L-Rhap-(1 → 3)-α-L-Rhap-(1 → OMe) (3). The semiflexible conformational behavior of these molecules was characterized by the occupation of a combination of different glycosidic linkage and side-chain conformational positions whose relative occupations were sensitive to dielectric screening. Molecular dynamics simulations of the trisaccharide 3 showed little difference between the linkage conformations in the trisaccharide and the component disaccharides 1 and 2. Experimental optical rotation data of 1 and 2 were obtained as a function of temperature in varying solvents. The molecular models were combined with the semiempirical theory of Stevens and Sathyanarayana to yield calculated optical rotations. Interpretation of the data of both 1 and 2 implied that a combination of conformations, both in glycosidic and side-chain positions, could explain the experimental data. Solvents effects were important in influencing the conformational mix and averaged optical rotation. Three-bond heteronuclear coupling constants 3JC, H were obtained for the glycosidic linkages of 1 and 2 in D2O and DMSO. Analysis of the coupling constants with a Karplus curve showed that small reductions in the glycosidic torsion angles of the conformations of the models used here of ca. 10°–15° in ϕ and 5°–10° in ψ were required to give better agreement with experiment; a combination of conformations for both 1 and 2 was consistent with the data. There was a negligible influence on the coupling constants of 1 on changing the solvent from D2O to DMSO. © 1997 John Wiley & Sons, Inc.  相似文献   

6.
What determines the shape of the allowed regions in the Ramachandran plot? Although Ramachandran explained these regions in terms of 1–4 hard‐sphere repulsions, there are discrepancies with the data where, in particular, the αR, αL, and β‐strand regions are diagonal. The αR‐region also varies along the α‐helix where it is constrained at the center and the amino terminus but diffuse at the carboxyl terminus. By analyzing a high‐resolution database of protein structures, we find that certain 1–4 hard‐sphere repulsions in the standard steric map of Ramachandran do not affect the statistical distributions. By ignoring these steric clashes (N···Hi+1 and Oi?1···C), we identify a revised set of steric clashes (Cβ···O, Oi?1···Ni+1, Cβ···Ni+1, Oi?1···Cβ, and Oi?1···O) that produce a better match with the data. We also find that the strictly forbidden region in the Ramachandran plot is excluded by multiple steric clashes, whereas the outlier region is excluded by only one significant steric clash. However, steric clashes alone do not account for the diagonal regions. Using electrostatics to analyze the conformational dependence of specific interatomic interactions, we find that the diagonal shape of the αR and αL‐regions also depends on the optimization of the N···Hi+1 and Oi?1···C interactions, and the diagonal β‐strand region is due to the alignment of the CO and NH dipoles. Finally, we reproduce the variation of the Ramachandran plot along the α‐helix in a simple model that uses only H‐bonding constraints. This allows us to rationalize the difference between the amino terminus and the carboxyl terminus of the α‐helix in terms of backbone entropy.  相似文献   

7.
Conformations of the α-l -Rhap(1-2)-β-d -Glc1-OMe and β-d -Galp(1-3)-β-d -Glc1-OMe disaccharides and the branched title trisaccharide were examined in DMSO-d6 solution by 1H-nmr. The distance mapping procedure was based on rotating frame nuclear Overhauser effect (NOE) constraints involving C- and O-linked protons, and hydrogen-bond constraints manifested by the splitting of the OH nmr signals for partially deuteriated samples. An “isotopomer-selected NOE” method for the unequivocal identification of mutually hydrogen-bonded hydroxyl groups was suggested. The length of hydrogen bonds thus detected is considered the only one motionally nonaveraged nmr-derived constraint. Molecular mechanics and molecular dynamics methods were used to model the conformational properties of the studied oligosaccharides. Complex conformational search, relying on a regular Φ,Ψ-grid based scanning of the conformational space of the selected glycosidic linkage, combined with simultaneous modeling of different allowed orientations of the pendant groups and the third, neighboring sugar residue, has been carried out. Energy minimizations were performed for each member of the Φ,Ψ grid generated set of conformations. Conformational clustering has been done to group the minimized conformations into families with similar values of glycosidic torsion angles. Several stable syn and anti conformations were found for the 1→2 and 1→3 bonds in the studied disaccharides. Vicinal glycosylation affected strongly the occupancy of conformational states in both branches of the title trisaccharide. The preferred conformational family of the trisaccharide (with average Φ,Ψ values of 38°, 17° for the 1→2 and 48°, 1° for the 1→3 bond, respectively) was shown by nmr to be stabilized by intramolecular hydrogen bonding between the nonbonded Rha and Gal residues. © 1998 John Wiley & Sons, Inc. Biopoly 46: 417–432, 1998  相似文献   

8.
A series of model compounds containing 3‐amino‐1H‐pyrazole‐5‐carboxylic acid residue with N‐terminal amide/urethane and C‐terminal amide/hydrazide/ester groups were investigated by using NMR, Fourier transform infrared, and single‐crystal X‐ray diffraction methods, additionally supported by theoretical calculations. The studies demonstrate that the most preferred is the extended conformation with torsion angles ? and ψ close to ±180°. The studied 1H‐pyrazole with N‐terminal amide/urethane and C‐terminal amide/hydrazide groups solely adopts this energetically favored conformation confirming rigidity of that structural motif. However, when the C‐terminal ester group is present, the second conformation with torsion angles ? and ψ close to ±180° and 0°, respectively, is accessible. The conformational equilibrium is observed in NMR and Fourier transform infrared studies in solution in polar environment as well as in the crystal structures of other related compounds. The observed conformational preferences are clearly related to the presence of intramolecular interactions formed within the studied residue. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
10.
Potential energy as contributions from non-bonded, electrostatic, hydrogen bonding and torsional interactions was computed as a function of dihedral angles around the glycosylic and exocyclic bonds for four important modified nucleic acid subunits, viz. (1) pseudouridine with unusual glycosylic bond, (2) dihydrouridine with saturated base ring, (3) N2-dimethyl guanosine having double methylation in the base ring, and (4) 2′-0-methyl adenosine having methylation in the ribose moiety. The two preferred, C2,-endo and C3,-endo, sugar puckers were considered. The probable low energy regions in the χ-ψ space and the population of various conformational states for each of the molecules were determined. The results of modified units were compared with those of the corresponding normal units. Experimental results available on simple molecular systems and on tRNA molecules were used for comparisons with theoretical predictions.  相似文献   

11.
For the development of a scheme for quantitative experimental estimation of internal motion in the complex human milk hexasaccharide lacto‐N‐di‐fuco hexose I (LNDFH I), we measured a large number of experimental residual dipolar couplings in liquid crystal orienting media. We present a total of 40 13C? 1H and 1H? 1H dipolar coupling values, each representing distinct directions of internuclear vectors. The NMR data were interpreted with established methods for analysis of rigid subdomains of the oligosaccharide as well as a novel method in which dipolar couplings were calculated over an ensemble of conformers from a solvent Molecular Dynamics trajectory using multiple linear regression analysis. The Lewisb epitope region of LNDFH I assumed a single unique conformation with internal motion described by fluctuations of 5–10° in glycosidic dihedral angles consistent with previous studies. Greater flexibility was observed for the remaining GlcNAc1→3‐β‐D ‐Gal and β‐D ‐Gal1→4Glc linkages, with the former glycosidic linkage existing in a conformational exchange among three states. The results were also supported by similar results of calculations carried out with conformers obtained from a simple Monte Carlo simulation without explicit solvent. © 2010 Wiley Periodicals, Inc. Biopolymers 95: 39–50, 2011.  相似文献   

12.
Two-dimensional nuclear magnetic resonance (n.m.r.) spectroscopy and a variety of computational techniques have been used to generate three-dimensional structures of the two DNA duplexes d(CGCCTAATCG) and d(CGTCACGCGC). The central six base-pairs in these two decamers contain all ten dinucleotide pairs in DNA and thus, represent a model system for investigating how the local structure of DNA varies with base sequence. Resonance assignments were made for the non-exchangeable base protons and most of the C-1'-C-4' sugar protons in both decamers. Three-dimensional structures were generated using a distance geometry algorithm and these initial structures were refined by optimizing the fit of back-calculated spectra against the experimental two-dimensional nuclear Overhauser effect (NOE) spectra. This back-calculation procedure consists of calculating NOE cross relaxation rates for a given structure by solution of the Bloch equations, and directly accounts for spin diffusion effects. Use of this refinement procedure eliminates some assumptions that have been invoked when generating structures of DNA oligomers from n.m.r. data. Constrained energy minimization and constrained quenched molecular dynamics calculation were also performed on both decamers to help generate energetically favorable structures consistent with the experimental data. Analysis of the local conformational parameters of helical twist, helical rise, propeller twist, displacement and the alpha, beta, gamma, epison and zeta backbone torsion angles in these structures shows that these parameters span a large range of values relative to the X-ray data of nucleic acids. However, the glycosidic and pseudorotation angles are quite well defined in these structures. The implications that these results have for determination of local structural variations of DNA in solution, such as those predicted by Callidine's rules, are discussed. Our results differ significantly from some previous studies on determining local conformations of nucleic acids and comparisons with these studies are made.  相似文献   

13.
The immobilization of the globular protein α‐1‐acid glycoprotein (AGP) onto silica gel led to the commercial availability of an AGP column, which has a high enantioselectivity. The enantioselectivity of AGP columns has been demonstrated in numerous applications. Due to potential AGP structural changes occurring upon its immobilization, the interaction between particular pairs of enantiomers and the stationary phase is very difficult to assess. Therefore, in this paper we report a mechanistic study that probes the nature of these types of interactions. As model ligands, we employed two LTD4 antagonists (L‐708, 738, MK0476, and their enantiomers) which have a rigid backbone consisting of a conjugated aromatic region and a side chain which is terminated with a carboxylic functional group. The difference between the two compounds is a two‐fluorine versus one‐chlorine substituent in the aromatic region of the molecule. To study the interaction between the two homologues and the AGP stationary phase, several parameters were varied, including pH, ionic strength, organic modifier, and temperature. van't Hoff plots were constructed and found to be nonlinear. They could, however, be divided into two linear regions, one from 0°C to ∼30°C, and another from 39°C to 50°C. The region at lower temperature implied that the separation was entropy‐dominated while the separation at higher temperature was enthalpically driven. The transition from the entropic to the enthalpically driven separation region suggested that bound AGP undergoes a conformational change. Fluorescence spectroscopy performed on the AGP stationary phase found evidence for a limited conformational transition at a similar temperature, consistent with this hypothesis. Chirality 11:224–232, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

14.
15.
A fundamental question in protein science is what is the intrinsic propensity for an amino acid to be in an α-helix, β-sheet, or other backbone dihedral angle (-ψ) conformation. This question has been hotly debated for many years because including all protein crystal structures from the protein database, increases the probabilities for α-helical structures, while experiments on small peptides observe that β-sheet-like conformations predominate. We perform molecular dynamics (MD) simulations of a hard-sphere model for Ala dipeptide mimetics that includes steric interactions between nonbonded atoms and bond length and angle constraints with the goal of evaluating the role of steric interactions in determining protein backbone conformational preferences. We find four key results. For the hard-sphere MD simulations, we show that (1) β-sheet structures are roughly three and half times more probable than α-helical structures, (2) transitions between α-helix and β-sheet structures only occur when the backbone bond angle τ (N–Cα–C) is greater than 110°, and (3) the probability distribution of τ for Ala conformations in the “bridge” region of-ψ space is shifted to larger angles compared to other regions. In contrast, (4) the distributions obtained from Amber and CHARMM MD simulations in the bridge regions are broader and have increased τ compared to those for hard sphere simulations and from high-resolution protein crystal structures. Our results emphasize the importance of hard-sphere interactions and local stereochemical constraints that yield strong correlations between -ψ conformations and τ.  相似文献   

16.
Three dimensional structures of sialyl Lewis(x) (SLe(x)) in aqueous solution and bound to selectinE are described based on an exhaustive conformational analysis and several long molecular dynamics simulations using different glycosidic regions as starting conformations. It appears from this study that when the oligosaccharide is free in solution the NeuNAcalpha(2-3)Gal segment favors glycosidic conformation in three different regions in the (Phi,Psi) plane with propensity of populations in the ratio 1:8:1. Each one of these structures is characteristically stabilized by specific hydrogen bonding interaction between NeuNAc and Gal. On the other hand, the Gal-GlcNAc-Fuc segment can exist in four different conformational states. Based on the topology of SLe(x) we are able to predict that out of all the allowed conformations in solution only two of these structures possess a geometry that would fit without steric clashes into the binding location of selectinE. In both of these binding modes, segment Gal-GlcNAc-Fuc adopts a unique conformation. The only difference between the two SLe(x) conformers that can successfully bind to selectinE is given by two possible regions in glycosidic space in the fragment NeuNAcalpha(2-3)Gal. A large conformational departure from the crystallographic data is observed for two lysine residues at the binding site of selectinE. These two residues play an important role when SLe(x) binds selectinE in aqueous solution. These findings help reconcile the X-ray data, in which these residues appear to be 1 nm away from SLe(x), with recent liquid NMR data reporting couplings between these protein residues and the sugar.  相似文献   

17.
The virtual bond scheme set forth in preceding papers for treating the average properties of polyriboadenylic acid (poly rA) is here applied to the calculation of the unperturbed mean-square end-to-end distance of polydeoxyriboadenylic acid (poly dA). The modifications in structure and in charge distribution resulting from the replacement of the hydroxyl group at C2′ in the ribose residue by hydrogen in deoxyribose produce only minor modifications in the conformational energies associated with the poly dA chain as compared to those found for poly rA. The main difference is manifested in the energy associated with rotations about the C3′–O3′ bond of the deoxyribose residue in the C2′-endo conformation; accessible rotations are confined to the range between 0° and 30° relative to the trans conformation, whereas in the ribose unit the accessible regions comprise two ranges centered at approximately 35° and 85°. The characteristic ratio 〈r2〉0/nl2 calculated on the basis of the conformational energy estimates is ≈9 for the poly dA chain with all deoxyribose residues in the C3′-endo conformation and ≈21 with all residues in the C2′-endo form. Satisfactory agreement is achieved between the theoretical values and experimental results on apurinic acid by treating the poly dA chain as a random copolymer of C3′-endo and C2′-endo conformational isomers present in a ratio of ~1 to 9.  相似文献   

18.
Voltage‐dependent anion channel 1 (VDAC1), which is located in the outer mitochondrial membrane, plays important roles in various cellular processes. For example, oligomerization of VDAC1 is involved in the release of cytochrome c to the cytoplasm, leading to apoptosis. However, it is unknown how VDAC1 oligomerization occurs in the membrane. In the present study, we determined high‐resolution crystal structures of oligomeric human VDAC1 (hVDAC1) prepared by using an Escherichia coli cell‐free protein synthesis system, which avoided the need for denaturation and refolding of the protein. Broad‐range screening using a bicelle crystallization method produced crystals in space groups C222 and P22121, which diffracted to a resolution of 3.10 and 3.15 Å, respectively. Each crystal contained two hVDAC1 protomers in the asymmetric unit. Dimer within the asymmetrical unit of the crystal in space group C222 were oriented parallel, whereas those of the crystal in space group P22121 were oriented anti‐parallel. From a model of the crystal in space group C222, which we constructed by using crystal symmetry operators, a heptameric structure with eight patterns of interaction between protomers, including hydrophobic interactions with β‐strands, hydrophilic interactions with loop regions, and protein–lipid interactions, was observed. It is possible that by having multiple patterns of interaction, VDAC1 can form homo‐ or hetero‐oligomers not only with other VDAC1 protomers but also with other proteins such as VDAC2, VDAC3 and apoptosis‐regulating proteins in the Bcl‐2 family.  相似文献   

19.
The yeast scaffold protein Pan1 contains two EH domains at its N‐terminus, a predicted coiled‐coil central region, and a C‐terminal proline‐rich domain. Pan1 is also predicted to contain regions of intrinsic disorder, characteristic of proteins that have many binding partners. In vitro biochemical data suggest that Pan1 exists as a dimer, and we have identified amino acids 705 to 848 as critical for this homotypic interaction. Tryptophan fluorescence was used to further characterize Pan1 conformational states. Pan1 contains four endogenous tryptophans, each in a distinct region of the protein: Trp312 and Trp642 are each in an EH domain, Trp957 is in the central region, and Trp1280 is a critical residue in the Arp2/3 activation domain. To examine the local environment of each of these tryptophans, three of the four tryptophans were mutagenized to phenylalanine to create four proteins, each with only one tryptophan residue. When quenched with acrylamide, these single tryptophan mutants appeared to undergo collisional quenching exclusively and were moderately accessible to the acrylamide molecule. Quenching with iodide or cesium, however, revealed different Stern‐Volmer constants due to unique electrostatic environments of the tryptophan residues. Time‐resolved fluorescence anisotropy data confirmed structural and disorder predictions of Pan1. Further experimentation to fully develop a model of Pan1 conformational dynamics will assist in a deeper understanding of the mechanisms of endocytosis. Proteins 2013; 81:1944–1963. © 2013 Wiley Periodicals, Inc.  相似文献   

20.
The disaccharide arabinobiose (5-O -α-L -arabinofuranosyl-α-L -arabinofuranose) constitutes the basic repeating structures found in such polysaccharides as arabinan or in the side chains of the hairy regions of pectins. The conformational behavior of aqueous arabinobiose has been investigated by high resolution nmr and computerized molecular modeling. The complete conformational analysis of the, disaccharide has been achieved with the MM3 molecular mechanics methods using the flexible residue method. In this study, both the puckering of the arabinofuranose, rings and the orientations about the glycosidic torsion angles ?, ψ, and ω; were considered. Some insights into conformational transitions were obtained through molecular dynamics simulation using the CHARMM force field. In parallel, transient nuclear Overhauser effects at 400.13 MHz and long-range vicinal homonuclear and heteronuclear coupling constants have been measured. The theoretical nmr data were calculated taking into account all accessible conformations and using averaging methods for both slow and fast internal motions models. The data do not support a single conformational model, and only conformational averaging yields the excellent agreement between the observed and simulated parameters. Within the potential energy surfaces computed for the disaccharide, several low energy conformers can be identified. When these conformations are extrapolated to regular polysaccharide structures, they generate chains of arabinan displaying right- and left-handed chirality and a wide range of repeating units per turn of helix. © 1994 John Wiley & Sons, Inc.  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