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.
Abstract

Adsorbed atomic monolayers of atoms such as carbon and nitrogen can cause substantial reconstructions of a nickel {001} surface. In this simulation we combine an atomic-orbital-based calculation of electronic structure with an empirical pair-wise repulsive potential to model the covalent part of the total energy. For 0.5 monolayer coverage by the adsorbate, the surface metal layer relaxes into a p(2 × 2) structure, with transverse displacements of about 0.4 Å. At the same time these displaced surface nickel atoms ride up above second layer nickels, with a vertical displcement of about 0.4 Å. The covalent contribution to the relaxation energy comes out at about 2.0 eV per carbon atom and 1.4eV per nitrogen atom, of which the reconstruction contributes about 0.3eV.  相似文献   

2.
Abstract

Although pore formation by protective antigen (PA) is critical to cell intoxication by anthrax toxin (AT), the structure of the pore form of PA (the PA63 pore) has not been determined. Hence, in this study, the PA63 pore was modeled using the X-ray structures of monomeric PA and heptameric α-hemolysin (α-HL) as templates. The PA63 pore model consists of two weakly associated domains, namely the cap and stem domains. The ring-like cap domain has a length of 80 Å and an outside diameter of 120 Å, while the cylinder-like stem domain has a length of 100 Å and outside diameter of ~28 Å. This provides the PA63 pore model with a length of 180 Å. Based on experimental results, the channel in the PA63 pore model was built to have a minimum diameter of ~12 Å, depending on side chain conformations. Because of its large size and structural complexity, the all-atom model of the PA63 pore is the end-stage construction of four separate modeling projects described herein. The final model is consistent with published experimental results, including mutational analysis and channel conductance experiments. In addition, the model was energetically and hydropathically refined to optimize molecular packing within the protomers and at the protomer-protomer interfaces. By providing atomic detail to biochemical and biophysical data, the PA63 pore model may afford new insights into the binding mode of PA on the membrane surface, the prepore-pore transition, and the mechanism of cell entry by anthrax toxin.  相似文献   

3.
Anecdotal and research evidence is that vertical jump performance declines over the competitive volleyball season. The purpose of this study was to evaluate whether a short period of ballistic resistance training would attenuate this loss. Fourteen collegiate women volleyball players were trained for 11 weeks with periodized traditional and ballistic resistance training. There was a 5.4% decrease (p < 0.05) in approach jump and reach height during the traditional training period (start of season to midseason), and a 5.3% increase (p < 0.05) during the ballistic training period (midseason to end of season), but values were not different from start to end of season. These changes in overall jump performance were reflective of changes in underlying neuromuscular performance variables: in particular, power output and peak velocity during loaded jump squats, countermovement jumps, and drop jumps. During the first 7 weeks of traditional heavy resistance training, it appears that the neuromuscular system is depressed, perhaps by the combination of training, game play, and skills practice precluding adequate recovery. Introduction of a novel training stimulus in the form of ballistic jump squats and reduction of heavy resistance training of the leg extensors stimulated a rebound in performance, in some cases to exceed the athlete's ability at the start of the season. Periodization of in-season training programs similar to that used in this study may provide volleyball players with good vertical jump performance for the crucial end-of-season games.  相似文献   

4.
Summary Watestriders (Gerris paludum F.), displaced by flowing water or wind, compensate for this by periodic jumps against the direction of drift so that they keep their average position — relative to the river bank, for instance — constant over long periods of time. To identify the cues used by the animals to compensate for drift, they were kept on an artificial stream with visual patterns along one or both sides. The velocity of the water flow and the pattern motion were varied. It is not possible to induce compensatory jumps in darkness by water or air current alone. Visual cues are indispensable for the reaction. The product of jump amplitude and jump frequency equals the drift velocity on average. The jump amplitudes are more or less independent of the flow velocity while the jump frequency is adjusted to it.  相似文献   

5.
Abstract

Thermodynamic and kinetic properties of the B-Z transition of poly(dG-m5dC) were investigated using polynucleotide samples ranging in length from 11000 to 300 base pairs. Van't Hoff enthalpy values increase with increasing polymer length for the B-Z transition in 0.35 mM MgCl2, 50 mM NaCl, 5 mM TRIS, pH 8. Rates of the B to Z transition increase with increasing polymer length for a jump of 0 to 3 mM MgCl, in 50 mM Nad, 5 mM TRIS, pH 8. The activation energy of the B to Z transition equals 7.9 ± 0.3 kcal/mol and is length independent Thermodynamic and kinetic data were fit to a model that simulates distribution of B- and Z-form tracts at the midpoint of B-Z equilibrium as a function of polymer length. A cooperative length of 1000 ± 200 base pairs is estimated for the B-Z transition. A direct relationship between rates of the B to Z transition and the square of the van't Hoff enthalpy values of the B-Z transition reflects a dependence of kinetics and cooperativity upon the energy of the nucleation event Faster B to Z transition rates with increasing polymer length can be explained by a mechanism rate limited by nucleation within the polymer instead of the ends.  相似文献   

6.
Carugo  Oliviero 《Amino acids》2020,52(3):435-443

A non-redundant set of 231 protein crystal structures refined at a resolution better than (or equal to) 1 Å was extracted from the Protein Data Bank and the degree of conformational rigidity at the protein-water interface was examined by means of the Hirshfeld test and by comparing the orientations of the anisotropic Us for contacting protein and water atoms. Contacts between protein and water atoms are more rigid that contacts between water atoms and the degree of rigidity increases for shorter contacts and for more hydrogen-bonded atoms. Nevertheless, water and protein atoms are not rigidly held together. On the contrary, they seem to have little influence on their mobility to such an extent that hydration water, different from the protein atoms, cannot be considered to be properly in the solid state

  相似文献   

7.
The structure is based on a difference Fourier synthesis at 2.8 Å resolution, using observed structure amplitudes and calculated phases, derived from a refinement of horse methaemoglobin at 2.0 Å resolution. Carbonmonoxyhaemoglobin has the same quaternary structure as methaemoglobin, but differs from it by slight changes in tertiary structure in the immediate vicinity of the haems. On transition from met- to carbonmonoxyhaemoglobin the distal histidines move away from the haem ligands towards the molecular surface, and both the haems and F-helices rotate slightly and shift towards the distal side. In methaemoglobin the sulphydryl group of cysteine F9(93)β is in equilibrium between two alternative positions: one external and the other half-buried in the “tyrosine pocket” between helices F and H. In carbonmonoxyhaemoglobin all the electron density for the sulphydryl group is in the half-buried position, so that the side chain of tyrosine HC2(145)β is completely displaced from its pocket. The difference map shows that the CO oxygen lies off the haem axis in both subunits, but the carbon cannot be seen as it coincides with the water molecule in methaemoglobin. A preliminary refinement of carbonmonoxyhaemoglobin suggests that the carbon may be displaced from the haem axis in the same direction as the oxygen. The haem pocket is so constructed that it fits an oxygen molecule in the bent conformation, but not a CO molecule which has its axis normal to the haem plane, because of steric hindrance by N? of the distal histidine and by Cγ2 of the distal valine. These two side chains apparently push the CO oxygen off the haem axis. The difference map indicates that in methaemoglobin the α-haem is ruffled and that on transition from met- to carbonmonoxyhaemoglobin it becomes flattened; in the β-haem the iron appears to move towards the porphyrin plane. The resolution is not sufficient to determine the exact position of the iron atoms and the proximal histidines relative to the porphyrins.  相似文献   

8.
Studies on the interactions between L ‐O‐ phosphoserine, as one of the simplest fragments of membrane components, and the Cinchona alkaloid cinchonine, in the crystalline state were performed. Cinchoninium L ‐O‐phosposerine salt dihydrate (PhSerCin) crystallizes in a monoclinic crystal system, space group P21, with unit cell parameters: a = 8.45400(10) Å, b = 7.17100(10) Å, c = 20.7760(4) Å, α = 90°, β = 98.7830(10)°, γ = 90°, Z = 2. The asymmetric unit consists of the cinchoninium cation linked by hydrogen bonds to a phosphoserine anion and two water molecules. Intermolecular hydrogen bonds connecting phosphoserine anions via water molecules form chains extended along the b axis. Two such chains symmetrically related by twofold screw axis create a “channel.” On both sides of this channel cinchonine cations are attached by hydrogen bonds in which the atoms N1, O12, and water molecules participate. This arrangement mimics the system of bilayer biological membrane. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

9.
Abstract

In this paper a coarse-grained method called elastic network interpolation (ENI) is used to generate feasible transition pathways between two given conformations of the core central domain of 16S Ribosomal RNA (16S rRNA). The two given conformations are the extremes generated by a molecular dynamics (MD) simulation, which differ from each other by 10Å in root-mean-square deviation (RMSD). It takes only several hours to build an ENI pathway on a 1.5GHz Pentium with 512 MB memory, while the MD takes several weeks on high-performance multi-processor servers such as the SGI ORIGIN 2000/2100. It is shown that multiple ENI pathways capture the essential anharmonic motions of millions of timesteps in a particular MD simulation. A coarse-grained normal mode analysis (NMA) is performed on each intermediate ENI conformation, and the lowest 1% of the normal modes (representing about 40 degrees of freedom (DOF)) are used to parameterize fluctuations. This combined ENI/NMA method captures all intermediate conformations in the MD run with 1.5Å RMSD on average. In addition, if we restrict attention to the time interval of the MD run between the two extreme conformations, the RMSD between the closest ENI/NMA pathway and the MD results is about 1Å. These results may serve as a paradigm for reducedDOF dynamic simulations of large biological macromolecules as well as a method for the reduced-parameter interpretation of massive amounts of MD data.  相似文献   

10.
The crystal and molecular structures of two α-aminoisobutyric acid (Aib)-containing diketopiperazines, cyclo(Aib-Aib) 1 and cyclo(Aib-L -Ile) 2 , are reported. Cyclo(Aib-Aib) crystallizes in the space group P1 with a = 5.649(3), b = 5.865(2), c = 8.363(1), α = 69.89(6), β = 113.04(8), γ = 116.0(3), and Z = 1, while 2 occurs in the space group P212121 with a = 6.177(1), b = 10.791(1), c = 16.676(1), and Z = 4. The structures of 1 and 2 have been refined to final R factors of 0.085 and 0.086, respectively. In both structures the diketopiperazine ring shows small but significant deviation from planarity. A very flat chair conformation is adopted by 1, in which the Cα atoms are displaced by 0.07 Å on each side of the mean plane, passing through the other four atoms of the ring. Cyclo(Aib-Ile) favors a slight boat conformation, with Aib Cα and Ile Cα atoms displaced by 0.11 and 0.05 Å on the same side of the mean plane formed by the other ring atoms. Structural features in these two molecules are compared with other related diketopiperazines.  相似文献   

11.
The main channel for H2O2 access to the heme cavity in large subunit catalases is twice as long as in small subunit catalases and is divided into two distinct parts. Like small subunit catalases, the 15 Å of the channel adjacent to the heme has a predominantly hydrophobic surface with only weak water occupancy, but the next 15 Å extending to the protein surface is hydrophilic and contains a complex water matrix in multiple passages. At the approximate junction of these two sections are a conserved serine and glutamate that are hydrogen bonded and associated with H2O2 in inactive variants. Mutation of these residues changed the dimensions of the channel, both enlarging and constricting it, and also changed the solvent occupancy in the hydrophobic, inner section of the main channel. Despite these structural changes and the prominent location of the residues in the channel, the variants exhibited less than a 2-fold change in the kcat and apparent KM kinetic constants. These results reflect the importance of the complex multi-passage structure of the main channel. Surprisingly, mutation of either the serine or glutamate to an aliphatic side chain interfered with heme oxidation to heme d.  相似文献   

12.
Abstract

This paper describes two complexes containing ethidium and the dinucleoside monophosphate, cytidylyl(3′-5′)guanosine (CpG). Both crystals are monoclinic, space group P21, with unit cell dimensions as follows: modification 1: a = 13.64 Å, b = 32.16 Å, c - 14.93 Å, β = 114.8° and modification 2: a = 13.79 Å, b = 31.94 Å, c = 15.66 Å, β = 117.5°. Each structure has been solved to atomic resolution and refined by Fourier and least squares methods; the first has been refined to a residual of 0.187 on 1,903 reflections, while the second has been refined to a residual of 0.187 on 1,001 reflections. The asymmetric unit in both structures contains two ethidium molecules and two CpG molecules; the first structure has 30 water molecules (a total of 158 non-hydrogen atoms), while the second structure has 19 water molecules (a total of 147 non-hydrogen atoms). Both structures demonstrate intercalation of ethidium between base-paired CpG dimers. In addition, ethidium molecules stack on either side of the intercalated duplex, being related by a unit cell translation along the a axis.

The basic feature of the sugar-phosphate chains accompanying ethidium intercalation in both structures is: C3′ endo (3′-5′) C2′ endo. This mixed sugar-puckering pattern has been observed in all previous studies of ethidium intercalation and is a feature common to other drug-nucleic acid structural studies carried out in our laboratory. We discuss this further in this paper and in the accompanying papers.  相似文献   

13.
The morphology of liver ribosomes and their subparticles, large and small, has been investigated. Analysis of the images has been carried out by successive selection of models and by X-raying them under conditions simulating negative staining. The relation between the main views has been checked by tilting the specimens in an electron microscope through ± 30 °.The small subparticle consists of an elongated body, to one of the ends of which a short “head” is attached. A model has been proposed, whose projections on rotation with respect to the longitudinal axis would satisfy all observable types of images. According to the proposed model, the “head” is tilted with respect to the elongated portion. The length of the subparticle is 230 Å. The dimensions of the elongated portion in the transverse direction are 110 Å × 75 to 80 Å.The large subparticles in lateral view resemble short “rods” 220 to 240 Å long and about 70 to 95 Å wide. At a distance of about 60 Å from the left end of the particles a projection (60 Å in length) is seen, on the inner side of which a depression, or “channel”, filled with the contrasting substance is always observed. Next to this depression a second projection is located, whose height is about 30 Å. The channel is either a cavity in the body of the large subparticle or a part of the RNA without protein. The length of the channel is about 80 Å, the width is about 50 to 60 Å. The left end of the particles is characterized by two sharpened portions; as a result, a cavity that shows an obtuse angle profile makes its appearance. The opposite end of the particles is cut off at an angle of 45 °. In another view, the subparticles appear to be almost rectangular in shape; they are characterized by dimensions of 150 Å × 220 to 240 Å. It is likely that the large projection is displaced sideways with respect to the longitudinal axis of the particles. The asymmetry associated with this displacement gives rise to preferred arrangements of the subparticles on the supporting film. An analysis has been made of the most typical images of monomeric ribosomes, on the basis of which a suggestion is made about mutual orientation of subparticles in a monomer.  相似文献   

14.
《Inorganica chimica acta》2001,312(1-2):226-230
A novel self-assembled polymer of Cu(TIM)+2 (where TIM=Me4[14]1,3,8,10-tetraeneN4, a macrocyclic ligand) with disulfate bridges has been synthesized and investigated by X-ray crystallography, IR, electronic and ESR spectroscopy. The compound crystallizes in the monoclinic space group C2/c with cell parameters a=16.745(4), b=7.4288(13), c=16.657(4) Å, β=115.789(19)°. The coordination geometry around the Cu(II) center is axially elongated distorted octahedral. The basal plane around the Cu center is planar with zero deviation from the weighted least square plane comprising four nitrogen atoms (N1, N2, N1′ and N2′) of the macrocyclic ligand. Two axial positions are occupied by the O atoms of the S2O7 2− molecules. The Cu⋯Cu distance is quite long (8.328 Å) along the chain. The X-band ESR spectra of a powdered solid sample of [CuTIM(S2O7)]n polymer, recorded at room temperature gives a isotropic spectrum with a g value of 2.22.  相似文献   

15.
The preparation, spectral properties, and crystal structure of a mononuclear copper(II) complex of acetylsalicylate and pyridine are reported. The complex exists as bis(acetylsalicylato)bis(pyridine)copper(II) both in the solid state and in chloroform solution. The crystal is monoclinic, space group P21/n, with a = 17.823(5), b = 10.903(4), c = 6.598(2) Å, β = 95.74(2)°. The final refinement used 1472 observed reflections and gave an R of 0.046. The copper atom is surrounded by four atoms in a trans square planar arrangement with two short CuO distances of 1.949(3) Å and two CuN distances of 2.003(4) Å. Two longer CuO distances of 2.623(3) Å are made with the remaining oxygen atoms of the aspirin carboxylate groups.  相似文献   

16.
A series of spin labels, varying in chain length between the maleimide attaching group and the nitroxide free radical, has been used to investigate the environment of the sulfhydryl group in human plasma albumin. From the electron spin resonance spectra, the degree of freedom of the nitroxide was determined and the location of the sulfhydryl was assessed. The effect of bound fatty acids on the sulfhydryl environment was also determined. The environment was found to be analogous to that in the bovine protein, that is, a crevice approximately 9.5 Å deep and not affected in the native state by fatty acids.  相似文献   

17.
Abstract

The crystal structure of 5′-amino-5′-deoxyadenosine (5′-Am.dA) p-toluenesulfonate has been determined by X-ray crystallographic methods. It belongs to the orthorhombic space group P212121 with a=7.754(3)Å, b=8.065(l)Å and c=32.481(2)Å. This nucleoside shows a syn conformation about the glycosyl bond and C2′-endo-C3′-exo puckering for the ribose sugar. The orientation of N5′ atom is gauche-trans about the exocyclic C4′-C5′ bond. The amino nitrogen N5′ forms a trifurcated hydrogen bond with N3, O9T and 04′ atoms. Adenine bases form A.A.A triplets through hydrogen bonding between N6, N7 and N1 atoms of symmetry related nucleoside molecules.  相似文献   

18.
《FEBS letters》1986,204(2):261-265
The effect of the conformational freedom of the ethanolamine tail of gramicidin A on the energy profile for the transfer of Na+, computed in the presence of water, shows an appreciable lowering of the minimum at 10.5 Å, and a splitting of the entrance barrier. The deep energy region at the channel mouth remains, however, the deepest one and contains a site of strong interaction of the ion with the Trp 11 carbonyl, at about 12 Å from the center.  相似文献   

19.
β‐Peptides are analogs of natural α‐peptides and form a variety of remarkably stable structures. Having an additional carbon atom in the backbone of each residue, their folded conformation is not only influenced by the side‐chain sequence but also and foremost by their substitution pattern. The precise mechanism by which the side chains interact with the backbone is, however, hitherto not completely known. To unravel the various effects by which the side chains influence the backbone conformation, we quantify to which extent the dihedral angles of a β3‐substited peptide with an additional methyl group on the central Cα‐atom can be regarded as independent degrees of freedom and analyze the distributions of these dihedral angles. We also selectively capture the steric effect of substituents on the Cα‐ and Cβ‐atoms of the central residue by alchemically changing them into dummy atoms, which have no nonbonded interactions. We find that the folded state of the β3‐peptide is primarily stabilized by a steric exclusion of large parts of the unfolded state (entropic effect) and only subsequently by mutual dependence of the ψ‐dihedral angles (enthalpic effect). The folded state of β‐peptides is stabilized by a different mechanism than that of α‐peptides. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

20.
Abstract

This paper describes two complexes containing N,N-dimethylproflavine and the dinucleoside monophosphate, 5-iodocytidylyl(3′-5′)guanosine (iodoCpG). The first complex is triclinic, space group PI, with unit cell dimensions a = 11.78 Å, b = 14.55 Å, c = 15.50 Å, a = 89.2°, β = 86.2°, γ = 96.4°. The second complex is monoclinic, space group P21, with a = 14.20 Å, b = 19.00 Å, c = 20.73 Å, β = 103.6°. Both structures have been solved to atomic resolution and refined by Fourier and least squares methods. The first structure has been refined anisotropically to a residual of 0.09 on 5,025 observed reflections using block diagonal least squares, while the second structure has been refined isotropically to a residual of 0.13 on 2,888 reflections with full matrix least squares. The asymmetric unit in both structures contains two dimethylproflavine molecules and two iodoCpG molecules; the first structure has 16 water molecules (a total of 134 non-hydrogen atoms), while the second structure has 18 water molecules (a total of 136 non-hydrogen atoms). Both structures demonstrate intercalation of dimethylproflavine between base-paired iodoCpG dimers. In addition, dimethylproflavine molecules stack on either side of the intercalated duplex, being related by a unit cell translation along b and a axes, respectively.

The basic structural feature of the sugar-phosphate chains accompanying dimethylproflavine intercalation in both structures is the mixed sugar puckering pattern: C3′ endo (3′-5′) C2′ endo. This same structural information is again demonstrated in the accompanying paper, which describes a complex containing dimethylproflavine with deoxyribo-CpG.

Similar information has already appeared for other “simple” intercalators such as ethidium, acridine orange, ellipticine, 9-aminoacridine, N-methyl-tetramethylphenanthrolinium and terpyridine platinum. “Complex” intercalators, however, such as proflavine and daunomycin, have given different structural information in model studies. We discuss the possible reasons for these differences in this paper and in the accompanying paper.  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