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 62 毫秒
1.
2.
Some endohedral fullerenes have been considered as possible candidate species for molecular memories. Recently, the encapsulation inside the fullerene cages has been extended from atoms to small molecules, for example the nitrogen molecule was placed inside the fullerene cage. The observed N2@C60 endohedral is computed in the paper together with NH3@C60, which was not yet observed. The computations are based on structural optimizations using density-functional theory (DFT) methods. In the optimized structures, the analytical harmonic vibrational analysis was carried out and the encapsulation energetics were evaluated using the second order Møller-Plesset (MP2) perturbation treatment. The lowest-energy structure has the N2 unit oriented towards a pair of parallel pentagons so that the complex exhibits D 5d symmetry. At the MP2 level, the encapsulation of N2 into C60 brings a potential energy gain of ? 9.3 kcal/mol while that for NH3 is ? 5.2 kcal/mol. The entropy term is also evaluated, yielding the standard Gibbs-energy change at room temperature for the encapsulation of N2 and NH3 of ? 2.6 and 1.5 kcal/mol, respectively. Some computed structural and vibrational characteristics are also reported. Emerging broader landscape of future applications of such encapsulates in nanoscience and nantechnology is discussed.  相似文献   

3.
Molecular dynamics (MD) computer simulations are utilized to better understand the dynamics of small (N = 5) endohedral Ne@C60 aggregates. Multiple runs at various temperatures are used to increase the reliability of our statistics. The aggregate holds together until somewhere between T = 1150 and 1200 K, where it dissociates, showing no intermediate sign of melting or fullerene disintegration. When the temperature is increased to around T = 4000 K, the encapsulated neon atoms begin to leave the aggregate, with the fullerene molecules still remaining intact. At temperatures near T = 4400 K, thermal disintegration of the fullerenes preempts the aggregate dissociation. Above this temperature neon atoms are more quickly released and the fullerenes form a larger connected structure, with bonding taking place in atom pairs from different original fullerene molecules. Escape constants and half lives are calculated for the temperature range 4000 K ≤ T ≤ 5000 K. The agreements and disagreements of results of this work with experiments suggest that classical MD simulations are useful in describing fullerene systems at low temperatures and near disintegration, but require development of new techniques before it is possible to accurately model windowing at temperatures below T = 3000 K.  相似文献   

4.
Molecular dynamics simulation is used to study the decomposition and stability of SII hydrogen and hydrogen/tetrahydrofuran (THF) hydrates at 150 K, 220 K and 100 bar. The modelling of the microscopic decomposition process of hydrogen hydrate indicates that the decomposition of hydrogen hydrate is led by the diffusive behaviour of H2 molecules. The hydrogen/THF hydrate presents higher stability, by comparing the distributions of the tetrahedral angle of H2O molecules, radial distribution functions of H2O molecules and mean square displacements or diffusion coefficients of H2O and H2 molecules in hydrogen hydrate with those in hydrogen/THF hydrate. It is also found that the resistance of the diffusion behaviour of H2O and H2 molecules can be enhanced by encaging THF molecules in the (51264) cavities. Additionally, the motion of THF molecules is restricted due to its high interaction energy barrier. Accordingly, THF, as a stabiliser, is helpful in increasing the stability of hydrogen hydrate.  相似文献   

5.
Fullerene molecules are cage-like nanoscopic structures with pentagonal and hexagonal faces. In practical applications such as fullerene-reinforced nanocomposites (FRNCs), these structures may be subjected to tension force. In this research, we employ molecular dynamics (MD) simulation to compute the behaviour and deformation of different fullerene molecules, ranging from C60 to C2000, under tension force. To model the interactions between carbon atoms in the MD simulations, the adaptive intermolecular reactive bond order (AIREBO) force field is used. The displacement–force and the displacement–strain energy curves are obtained. It is observed that a new type of structural instability occurs in the fullerene molecules when the applied tension force increases. This abnormal structural instability in the fullerenes is investigated for the first time in the literature. The critical tensile forces and the corresponding mode shapes are determined for different fullerenes. The results indicate that the critical forces and deformations strongly depend upon the number of carbon atoms.  相似文献   

6.
The influence of tacticity on chain dimensions, backbone and side-group conformational states and relaxation dynamics, intermolecular hydrogen bonding and its relaxation dynamics was investigated for 30 repeat unit poly(methacrylic acid) (PMA) as a function of the degree-of-neutralisation, f (i.e. charge density) [0 < f < 1 range] in explicit water with Na+ neutralising counter-ions, using molecular dynamics simulations. Chain expansion with increase in charge density is observed. Simulation results, where applicable, compare well with experimental results in the literature. Isotactic (i-PMA) chain exhibits the largest expansion (89% change in Rg), and syndiotactic (s-PMA-RR-0.7) chain shows 31%. For fully neutralised chain (high charge density), the probability for trans conformation at backbone bonds follows the trend i-PMA > a-PMA > s-PMA. At high charge density, a higher probability of trans state is obtained at meso dyads as compared to racemic dyads and the side group in i-PMA relative to other tacticity is conformationally more extended. RR triads impart better hydrogen bonding with water. In the COOH side group, greater rotational mobility is observed for i-PMA as compared to s-PMA. Water coordination to PMA is invariant with tacticity at low f. For f = 1, s-PMA exhibits the best level of H-bonding, due to its 3D chemical structural configuration. Binding distance between water and PMA atoms and their coordination number remain unchanged with respect to ionisation of COOH groups. Isotactic (i-PMA) exhibits the fastest relaxation of polymer-water H-bonds. While counter-ion coordination on to PMA is unaffected by tacticity, the intensity of ion condensation as well as hydrogen bond relaxation time at f = 1 varies in the order i-PMA > a-PMA > s-PMA.  相似文献   

7.
The microsolvation of aminoethanol (AE) with one, two, three or four water molecules was investigated using a density functional theory (DFT) approach. Quantum theory of atoms in molecules (QTAIM) analyses were employed to elucidate the hydrogen-bonding characteristics of AE–(H2O) n (n = 1–4) complexes. The results showed that AE tends to break its intramolecular OHAE···NAE hydrogen bond (H-bond) upon microsolvation and form intermolecular H-bonds with water molecules, while complexes that retain the intramolecular OHAE···NAE H-bond show reduced stabilities. The intermolecular H-bond that forms between the nitrogen atom of AE and the hydroxyl of a water molecule is the strongest one for the most stable AE–(H2O) n (n = 1–4) complexes, and as n increases from 1 to 4 they grow stronger. The partial covalent character of this H-bond was confirmed by QTAIM analyses. Many-body interaction analysis showed that the relaxation energies and two- and three-body energies make significant contributions to the binding energies of the complexes.  相似文献   

8.
The kinetics of the coil-to-helix transition of (dG-dC)3 in M NaCl, 45 mM sodium cacodylate, pH 7, were measured in H2O, D2O, 10 mol % ethanol, 10 mol % urea, and 10 mol % glycerol. At 43°C in H2O the recombination rate is 1.3 ± 0.2 × 107 M?1 s?1; the dissociation rate is 68 ± 10 s?1. The destabilization of the helix in 10 mol % ethanol and 10 mol % urea relative to water is primarily due to a large increase in the helix-dissociation rate. In 10 mol % glycerol, the destabilization of the helix is due to a decrease in the recombination rate and an increase in the dissociation rate. Above 20°C, two exponential decays longer than 1 μs are observed after a temperature jump. The slower relaxation time is 4–10 times faster than the bimolecular component and is independent of oligomer concentration. We attribute this relaxation to a rapid equilibrium between two helical states. At low temperatures and oligomer concentrations of 1 mM or greater, the helices aggregate in 1M NaCl. Experimental data are presented under conditions where aggregation is unimportant and evidence is given that the ΔH-determined spectroscopically is unaffected by aggregation.  相似文献   

9.
The adsorption behaviour of gas molecules on detector surfaces has a profound influence on the sensitivity of the detector. For this reason, this study used molecular dynamics simulation to explore the dynamic adsorption behaviour of hydrogen sulphide (H2S) molecules on various types of Au surfaces, including a planar Au(1?1?0) structure and three types of slit array structures. The influence of system temperature, adsorbate concentration and the slit width of nanoarrays on diffusivity, average adsorption energy and static adsorption amount were systematically examined. Simulation results indicate that the self-diffusivity of the adsorbate molecules increases with temperature but decreases with adsorbate concentration. At low concentrations (~3 mol/L), each type of Au(1?1?0) surface structure shows good capacity to adsorb all H2S molecules. With increasing concentration at 6.5 mol/L, the high concentration leads to adsorption saturation and many free H2S molecules in the planar Au(1?1?0) structure. Moreover, desorption also begins to appear on the planar structures at a temperature of 300 K (at 6.5 mol/L). The simulation results indicate that the columnar array structures with a slit width ≥5.76 Å allow molecules to swiftly spread into the slits and provide more stable adsorption sites (i.e. with a higher adsorption energy), which can effectively address the issues of high-temperature desorption and adsorption saturation. Particularly at low temperatures (≤100 K), slit structures presented a level of static adsorption of H2S that was 30% to 35 higher than that of planar structures.  相似文献   

10.
Terahertz absorption spectra of alanine polypeptides in water were simulated with classical molecular dynamics at 310 K. Vibrational modes and oscillator strengths were calculated based on a quasi-harmonic approximation. Absorption spectra of Alan (n = 5, 15, 30) with different chain lengths and Ala15 in coiled and helical conformations were studied in 10–40 cm? 1 bandwidth. Simulation results indicated both the chain length and the conformation have significant influences on THz spectra of alanine polypeptides. With the increase of chain length, the average THz absorption intensity increases. Compared with the helical Ala15 polypeptide, the THz spectra of coiled one shows stronger absorption peaks. These results were explained from different numbers of hydrogen bonds formed between polypeptides and the surrounding water molecules.  相似文献   

11.
Coumarin molecules have biological activities possessing lipid-controlling activity, anti-hepatitis C activity, anti-diabetic, anti-Parkinson activity, and anti-cancer activity. Here, we have presented an inclusive study on the interaction of 8-substituted-7-hydroxy coumarin derivatives (Umb-1/Umb-2) with α-1-glycoprotein (AGP) and human serum albumin (HSA) which are the major carrier proteins in the human blood plasma. Binding constants obtained from fluorescence emission data were found to be KUmb-1=3.1 ± .01 × 104 M?1, KUmb-2 = 7 ± .01 × 104 M?1, which corresponds to ?6.1 and ?6.5 kcal/mol of free energy for Umb-1 and Umb-2, respectively, suggesting that these derivatives bind strongly to HSA. Also these molecules bind to AGP with binding constants of KUmb-1-AGP=3.1 ± .01 × 103 M?1 and KUmb-2-AGP = 4.6 ± .01 × 103 M?1. Further, the distance, r between the donor (HSA) and acceptor (Umb-1/Umb-2) was calculated based on the Forster’s theory of non-radiation energy transfer and the values were observed to be 1.14 and 1.29 nm in Umb-1–HSA and Umb-2–HSA system, respectively. The protein secondary structure of HSA was partially unfolded upon binding of Umb-1 and Umb-2. Furthermore, site displacement experiments with lidocaine, phenylbutazone (IIA), and ibuprofen (IIIA) proves that Umb derivatives significantly bind to subdomain IIIA of HSA which is further supported by docking studies. Furthermore, Umb-1 binds to LYS402 with one hydrogen bond distance of 2.8 Å and Umb-2 binds to GLU354 with one hydrogen bond at a distance of 2.0 Å. Moreover, these molecules are stabilized by hydrophobic interactions and hydrogen bond between the hydroxyl groups of carbon-3 of coumarin derivatives.  相似文献   

12.
FtsZ is an appealing target for the design of antimicrobial agent that can be used to defeat the multidrug-resistant bacterial pathogens. Pharmacophore modelling, molecular docking and molecular dynamics (MD) simulation studies were performed on a series of three-substituted benzamide derivatives. In the present study a five-featured pharmacophore model with one hydrogen bond acceptors, one hydrogen bond donors, one hydrophobic and two aromatic rings was developed using 97 molecules having MIC values ranging from .07 to 957 μM. A statistically significant 3D-QSAR model was obtained using this pharmacophore hypothesis with a good correlation coefficient (R2 = .8319), cross validated coefficient (Q2 = .6213) and a high Fisher ratio (F = 103.9) with three component PLS factor. A good correlation between experimental and predicted activity of the training (R2 = .83) and test set (R2 = .67) molecules were displayed by ADHRR.1682 model. The generated model was further validated by enrichment studies using the decoy test and MAE-based criteria to measure the efficiency of the model. The docking studies of all selected inhibitors in the active site of FtsZ protein showed crucial hydrogen bond interactions with Val 207, Asn 263, Leu 209, Gly 205 and Asn-299 residues. The binding free energies of these inhibitors were calculated by the molecular mechanics/generalized born surface area VSGB 2.0 method. Finally, a 15 ns MD simulation was done to confirm the stability of the 4DXD–ligand complex. On a wider scope, the prospect of present work provides insight in designing molecules with better selective FtsZ inhibitory potential.  相似文献   

13.
The influences of water solvent on the structures and stabilities of the complex ion conformers formed by the coordination of alanine dipeptide (AD) and Na+ have been investigated using supramolecular and polarizable continuum solvation models at the level of B3LYP/6-311++G**, respectively; 12 monohydrated and 12 dihydrated structures of Na+–AD complex ion were obtained after full geometrical optimization. The results showed that H2O molecules easily bind with Na+ of Na+–AD complex ion, forming an ion-lone pair interaction with the Na–O bond length of 2.1–2.3 Å. Besides, H2O molecules also can form hydrogen bonds OW–HW···O(1), OW–HW···O(2), N(1)–H(1)···OW or N(2)–H(2)···OW with O or N groups of the Na+–AD backbone. The most stable gaseous bidentate conformer C7AB of Na+–AD is still the most stable one in the solvent of water. However, the structure of the most unstable gaseous conformer α′B of Na+–AD collapses under the attack of H2O molecules and changes into C7AB conformation. Computations with IEFPCM solvation model of self-consistent reaction field theory give that aqueous C5A is more stable than C7eqB and that the stabilization energies of water solvent on monodentate conformers of Na+–AD complex ion (about 272–294 kJ/mol) are more than those on bidentate ones (about 243 kJ/mol).  相似文献   

14.
Density functional theory BLYP/DNP was employed to optimize a series of fullerenes and their holmium endohedral compounds, including C20, Ho@C20, Ho3+@C20, C60, Ho@C60, Ho3+@C60,C70, Ho@C70, Ho3+@C70 C78, Ho@C78, Ho3+@C78, C82,Ho@C82 and Ho3+@C82. DFT semi core pseudospot approximation was taken into consideration in the calculations of the element holmium because of its particular electronic structure. Fullerenes and their holmium endohedral compounds’ aromaticity were studied in terms of structural criteria, energetic criteria, and reactivity criteria. The results indicate that the aromaticity of fullerenes was reduced when a holmium atom was introduced into the carbon cage, and the endohedral fullerenes’ reactive activity enhance; but the aromaticity of the carbon cage increased when a Ho3+ cation was encapsulated into a fullerene. Calculations of aromaticity and stability indicate that two paths can lead to the similar aim of preparing holmium endohedral fullerenes; that is, they can form from either a holmium atom or a holmium cation (Ho3+) reacting with fullerenes, respectively, and the latter is more favorable.  相似文献   

15.
The structure of the peptide Boc-Val-Ala-Leu-Aib-Val-Ala-Leu-OMe has been determined in crystals obtained from a dimethylsulfoxide–isopropanol mixture. Crystal parameters are as follows: C38H69N7O10 · H2O · 2C3H7OH, space group P21, a = 10.350 (2) Å, b = 26.084 (4) Å, c = 10.395(2) Å, β = 96.87(12), Z = 2, R = 8.7% for 2686 reflections observed > 3.0 σ (F). A single 5 → 1 hydrogen bond is observed at the N-terminus, while two 4 → 1 hydrogen bonds characteristic of a 310-helix are seen in the central segment. The C-terminus residues, Ala(6) and Leu(7) are expended, while Val(5) is considerably distorted from a helical conformation. Two isopropanol molecules make hydrogen bonds to the C-terminal segment, while a water molecule interacts with the N-terminus. The structure is in contrast to that obtained for the same peptide in crystals from methanol-water [ I. L. Karle, J. L. Flippen-Anderson, K. Uma, and P. Balaram (1990) Proteins: Structure, Function and Genetics, Vol. 7, pp. 62–73] in which two independent molecules reveal an almost perfect α-helix and a helix penetrated by a water molecule. A comparison of the three structures provides a snapshot of the progressive effects of solvation leading to helix unwinding. The fragility of the heptapeptide helix in solution is demonstrated by nmr studies in CDC13 and (CD3)2SO. A helical conformation is supported in the apolar solvent CDCl3, whereas almost complete unfolding is observed in the strongly solvating medium (CD3)2SO. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
Crystals of 5‐fluorouridine (5FUrd) have unit cell dimensions a = 7.716(1), b = 5.861(2), c = 13.041(1)Å, α = γ = 90°, β = 96.70° (1), space group P21, Z = 2, ρobs = 1.56 gm/c.c and ρcalc = 1574 gm/c.c The crystal structure was determined with diffractometric data and refined to a final reliability index of 0.042 for the observed 2205 reflections (I ≥ 3σ). The nucleoside has the anti conformation [χ = 53.1(4)°] with the furanose ring in the favorite C2′–endo conformation. The conformation across the sugar exocyclic bond is g+, with values of 49.1(4) and ? 69.3(4)° for Φθc and Φ respectively. The pseudorotational amplitude τm is 34.5 (2) with a phase angle of 171.6(4)°. The crystal structure is stabilized by a network of N–H…O and O–H…O involving the N3 of the uracil base and the sugar O3′ and O2′ as donors and the O2 and O4 of the uracil base and O3′ oxygen as acceptors respectively. Fluorine is neither involved in the hydrogen bonding nor in the stacking interactions. Our studies of several 5‐fluorinated nucleosides show the following preferred conformational features: 1) the most favored anti conformation for the nucleoside [χ varies from ? 20 to + 60°] 2) an inverse correlation between the glycosyl bond distance and the χ angle 3) a wide variation of conformations of the sugar ranging froni C2′–endo through C3′–endo to C4′–exo 4) the preferred g+ across the exocyclic C4′–C5′ bond and 5) no role for the fluorine atom in the hydrogen bonding or base stacking interactions.  相似文献   

17.
The interaction between puerarin and β-cyclodextrin (CD) has been studied in D2O, H2O/acetone-d 6, acetone-d 6 and DMSO-d 6 solutions by 1H NMR spectroscopy. The NMR data obtained from hydroxy protons indicate that the formation of the inclusion complex between the two molecules is not stabilized by strong hydrogen bond interactions. The sugar part of puerarin as well as the A ring are outside the β-CD cavity while the B and C rings are located inside the cavity and the interaction is mainly stabilized by hydrophobic interactions. In DMSO at 30°C and in acetone-d 6/H2O at temperature below −5°C, doubling of some signals indicated that, in these solvent systems, free rotation of the C-glycosyl bond was restricted due to the steric hindrance between the phenolic hydroxy group at C-7 and the bulky sugar moiety at C-8. In acetone, fast exchange of phenolic protons on the NMR timescale was observed, showing the effect of the solvent on the hindered rotation.  相似文献   

18.
The structural stability of 8 × ([D-Leu-L-Lys-(D-Gln-L-Ala)3]) cyclic peptide nanotube (CPN) in water and different phospholipid bilayers were explored by 100 ns independent molecular dynamics (MD) simulations. The role of non-bonded interaction energy between the side and main chains of cyclic peptide rings in different membrane environments assessed, wherein the repulsive electrostatic interaction energy between neighbouring cyclic peptide rings was found adequate to break hydrogen bond energy thereby to crumple CPN. Further, the water permeation across the CPN channel was studied in four types of phospholipid bilayers- DMPG (1,2-Dimyristoyl-sn-glycero-3-phosphorylglycerol), DMPS (1,2-Dimyristoyl-sn-glycero-3-phosphoserine), POPC (1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine) and POPE (1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine) from MD simulations. DMPS membrane shows higher non-bonded interaction energies (?1913.06 kJ/mol of electrostatic interaction energy and ?994.13 kJ/mol of van der Waals interaction energy) with CPN due to the presence of polar molecules in lipid structure. Thusly, the non-bonded interaction energies were essential towards the stability of CPN than hydrogen bonds between the nearby cyclic peptides. The result also reveals the role of side chains, hydrogen bonds and non-bonded interaction energies in an aqueous environment. The diffusion coefficient of water obtained from means square deviation calculation shows similar coefficients irrespective of the lipid surroundings. However, the permeation coefficients demonstrate water flow in the channel relies upon the environment.  相似文献   

19.
Based on Hamid model of 11Å tobermorite, amorphous calcium silicates hydrates (or C-S-H) structures (Ca4Si6O14(OH)4?2H2O as the C-S-H(I) and (CaO)1.67(SiO2)(H2O)1.75 as the C-S-H(II)) with the Ca/Si ratio of 0.67 and 1.7 are concerned. Then, as the representative ‘globule’ C-S-H, two amorphous C-S-H structures with the size of 5.352 × 4.434 × 4.556 nm3 during the stretch process are simulated at a certain strain rate of 10?3 ps?1 by LAMMPS program for molecular dynamics simulation, using ClayFF force field. The tensile stress–strain curves are obtained and analysed. Besides, elastic modulus of the ‘globule’ C-S-H is calculated to assess the elastic modulus of C-S-H phases (the low-density C-S-H – LD C-S-H – and the high-density C-S-H – HD C-S-H), where the porosity is a critical factor for explaining the relationship between ‘globule’ C-S-H at nanoscale and C-S-H phases at microscale. Results show that: (1) The C-S-H(I) structure has transformed from crystalline to amorphous during the annealing process, Young’s moduli in x, y and z directions are almost the same. Besides, the extent of aggregation and aggregation path for water molecules in the structure is different in three directions. (2) Young’s modulus of both amorphous C-S-H(I) and C-S-H(II) structures with a size of about 5 nm under strain rate of 10?3 ps?1 at 300 K in three directions is averaged to be equal, of which C-S-H(II) structure is about 60.95 GPa thus can be seen as the elastic modulus of the ‘globule’ C-S-H. (3) Based on the ‘globule’ C-S-H, the LD C-S-H and HD C-S-H can be assessed by using the Self-Consistent Scheme (separately 18.11 and 31.45 GPa) and using the Mori–Tanaka scheme (29.78 and 37.71 GPa), which are close to the nanoindentation experiments by Constantinides et al. (21.7 and 29.4 GPa).  相似文献   

20.
In this study, we perform grand canonical Monte Carlo simulations to evaluate the hydrogen storage capacity of structure I (sI) hydrogen hydrates at pressures up to 500 MPa. Initially, we calculate the upper limit of H2 content of sI hydrates by studying the hypothetical sI hydrate, where H2 is the single guest component. It is found that the storage capacity of the hypothetical pure H2 sI hydrate could reach 3.5 wt% at 500 MPa and 274 K. Depending on pressure, the large cavities of the pure H2 hydrate can accommodate up to three H2 molecules while the small ones are singly occupied at most, even at pressures as high as 500 MPa, without any double occupancy being observed. Subsequently, the binary H2–ethylene oxide (EO) hydrate is examined. In this case, the large cavities are occupied by a single EO molecule while the small cavities can accommodate at most a single H2 molecule. Such configuration results in a maximum H2 content of only 0.37 wt%. The hydrogen storage capacity does not improve significantly even in case when EO is replaced by a component with smaller molecular weight.  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