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.
We have investigated the adsorption properties of acetone on zigzag single-walled BNNTs using density functional theory (DFT) calculations. The results obtained show that acetone is strongly bound to the outer surface of a (5,0) BNNT on the top site directly above the boron atom, with a binding energy of ?96.16 kJ?mol?1 and a B–O binding distance of 1.654 Å. Our first-principles calculations also predict that the ability of zigzag BNNTs to adsorb acetone is significantly stronger than the corresponding ability of zigzag CNTs. A comparative investigation of BNNTs with different diameters indicated that the ability of the side walls of the tubes to adsorb acetone decreases significantly for nanotubes with larger diameters. Furthermore, the stability of the most stable acetone/BNNT complex was tested using ab initio molecular dynamics simulation at room temperature.
Figure
First-principles calculations predict that acetone is strongly bound to the outer surfaces of BNNTs with a binding energy of ?107.14 kJ?mol?1. Comparison with the corresponding adsorption on CNTs reveals that the ability of BNNTs to adsorb acetone is about threefold that of CNTs.  相似文献   

2.
Covalent organic frameworks (COFs) are a class of covalently linked crystalline nanoporous materials, versatile for nanoelectronic and storage applications. 3D COFs, in particular, have very large pores and low mass densities. Extensive theoretical studies of their energetic and mechanical stability, as well as their electronic properties, have been carried out for all known 3D COFs. COFs are energetically stable and their bulk modulus ranges from 3 to 20 GPa. Electronically, all COFs are semiconductors with band gaps corresponding to the HOMO–LUMO gaps of the building units.
Figure
 3D covalent organic frameworks  相似文献   

3.
The structural, elastic and electronic properties of lutatium-pnictides (LuN, LuP, LuAs, LuSb, and LuBi) were analyzed by using full-potential linearized augmented plane wave within generalized gradient approximation in the stable rock-salt structure (B1 phase) with space group Fm-3m and high-pressure CsCl structure (B2 phase) with space group Pm-3m. Hubbard-U and spin-orbit coupling were included to predict correctly the semiconducting band gap of LuN. Under compression, these materials undergo first-order structural transitions from B1 to B2 phases at 241, 98, 56.82, 25.2 and 32.3 GPa, respectively. The computed elastic properties show that LuBi is ductile by nature. The electronic structure calculations show that LuN is semiconductor at ambient conditions with an indirect band gap of 1.55 eV while other Lu-pnictides are metallic. It was observed that LuN shows metallization at high pressures. The structural properties, viz, equilibrium lattice constant, bulk modulus and its pressure derivative, transition pressure, equation of state, volume collapse, band gap and elastic moduli, show good agreement with available data.
Figure
Equation of state of Lu-pnictides  相似文献   

4.
5.
To understand the chemical behavior of uranyl complexes in water, a bis-uranyl [(phen)(UO2)(μ2–F)(F)]2 (A; phen?=?phenanthroline, μ2?=?doubly bridged) and its hydrated form A?·?(H2O)n (n?=?2, 4 and 6) were examined using scalar relativistic density functional theory. The addition of water caused the phen ligands to deviate slightly from the U22–F)2 plane, and red-shifts the U–F-terminal and U?=?O stretching vibrations. Four types of hydrogen bonds are present in the optimized hydrated A?·?(H2O)n complexes; their energies were calculated to fall within the range 4.37–6.77 kcal mol-1, comparable to the typical values of 5.0 kcal mol-1 reported for hydrogen bonds. An aqueous environment simulated by explicit and/or implicit models lowers and re-arranges the orbitals of the bis-uranyl complex.
Figure
A bis(uranyl) complex [(phen)(UO2)(μ2–F)(F)]2 (A) and its solvated form A?·?(H2O)n were examined using scalar relativistic density functional theory. Hydrogen bonds cause the phen ligand to slightly deviate from the equatorial plane of the uranyl ion, resulting in a pronounced red-shift of the U–F-terminal and U?=?O asymmetric stretching vibrations. The calculated energies fall within 4.4?–6.8 kcal/mol. Explicit and/or implicit aqueous solvation re-arranges the molecular orbitals of the complex  相似文献   

6.
Quantum chemical calculations were performed for LiNH2–HMgX (X?=?H, F, Cl, Br, CH3, OH, and NH2) complexes to propose a new interaction mechanism between them. This theoretical survey showed that the complexes are stabilized through the combinative interaction of magnesium and lithium bonds. The binding energies are in the range of 63.2–66.5 kcal mol?1, i.e., much larger than that of the lithium bond. Upon complexation, both Mg–H and Li–N bonds are lengthened. Substituents increase Mg-H bond elongation and at the same time decrease Li-N bond elongation. These cyclic complexes were characterized with the presence of a ring critical point and natural population analysis charges.
Figure
A new interaction mechanism has been suggested for the LiNH2-HMgH complex. It was found that the combinative interaction of magnesium and lithium bonds is responsible for the stability of the complex. The effect of subsitutents on its stability has also been investigated  相似文献   

7.
The interactions of L-aminoglucosidic stereoisomers such as rhodostreptomycins A (Rho A) and B (Rho B) with cations (Mg2+, Ca2+, and H+) were studied by a quantum mechanical method that utilized DFT with B3LYP/6-311G**. Docking studies were also carried out in order to explore the surface recognition properties of L-aminoglucoside with respect to Mg2+ and Ca2+ ions under solvated and nonsolvated conditions. Although both of the stereoisomers possess similar physicochemical/antibiotic properties against Helicobacter pylori, the thermochemical values for these complexes showed that its high affinity for Mg2+ cations caused the hydration of Rho B. According to the results of the calculations, for Rho A–Ca2+(H2O)6, ΔH = ?72.21 kcal?mol?1; for Rho B–Ca2+(H2O)6, ΔH = ?72.53 kcal?mol?1; for Rho A–Mg2+(H2O)6, ΔH = ?72.99  kcal?mol?1 and for Rho B–Mg2+(H2O)6, ΔH = ?95.00  kcal?mol?1, confirming that Rho B binds most strongly with hydrated Mg2+, considering the energy associated with this binding process. This result suggests that Rho B forms a more stable complex than its isomer does with magnesium ion. Docking results show that both of these rhodostreptomycin molecules bind to solvated Ca2+ or Mg2+ through hydrogen bonding. Finally, Rho B is more stable than Rho A when protonation occurs.
Figure
Rho B–H showed higher stability since it is considered a proton pump inhibitor, and is therefore a stronger inhibitor of Helicobacter pylori  相似文献   

8.
DFT calculations, with VWN exchange correlation functional and double numeric basis set, were used to evaluate the energies required for the scission reactions taking place in the initial stage of the thermal degradation of Poly(methyl methacrylate) (PMMA) in the presence of a carbon nanotube (CNT). Side group and main chain scissions were investigated. The results averaged from five configurations of pure PMMA (DP?=?5) were used as references and compared to the results obtained for the five same configurations of PMMA grafted on three carbon nanotubes of similar diameter (1.49 nm). The bond dissociation energies (BDE) of main chain scission evaluated for grafted PMMA was 4 % less endothermic than for pure PMMA. These results seemed independent of the tested chirality (11,11); (12,10) and (16,5) of the carbon nanotubes. Comparisons with the BDE of the weakest bonds due to intrinsic defaults (head to head and unsaturated end chain) were performed.
Figure
Representation of PMMA radical oligomer and grafted carbon nanotube  相似文献   

9.
Increase of the atmospheric concentration of halogenated organic compounds is partially responsible for a change of the global climate. In this work we have investigated the interaction between halogenated ether and water, which is one of the most important constituent of the atmosphere. The structures of the complexes formed by the two most stable conformers of enflurane (a volatile anaesthetic) with one and two water molecules were calculated by means of the counterpoise CP-corrected gradient optimization at the MP2/6–311++G(d,p) level. In these complexes the CH…Ow hydrogen bonds are formed, with the H…Ow distances varying between 2.23 and 2.32 Å. A small contraction of the CH bonds and the blue shifts of the ν(CH) stretching vibrations are predicted. There is also a weak interaction between one of the F atoms and the H atom of water, with the Hw…F distances between 2.41 and 2.87 Å. The CCSD(T)/CBS calculated stabilization energies in these complexes are between ?5.89 and ?4.66 kcal?mol?1, while the enthalpies of formation are between ?4.35 and ?3.22 kcal?mol?1. The Cl halogen bonding between enflurane and water has been found in two complexes. The intermolecular (Cl···O) distance is smaller than the sum of the corresponding van der Waals radii. The CCSD(T)/CBS stabilization energies for these complexes are about ?2 kcal?mol?1.
Figure
Complex between enflurane and water molecules  相似文献   

10.
We studied hydrated calcium oxalate and its ions at the restricted Hartree–Fock RHF/6-31G* level of theory. Performing a configurational search seems to improve the fit of the HF/6-31G* level to experimental data. The first solvation shell of calcium oxalate contains 13 water molecules, while the first solvation shell of oxalate ion is formed by 14 water molecules. The first solvation shell of Ca(II) is formed by six water molecules, while the second shell contains five. At 298.15 K, we estimate the asymptotic limits (infinite dilution) of the total standard enthalpies of hydration for Ca(II), oxalate ion and calcium oxalate as ?480.78, –302.78 and –312.73 kcal mol?1, resp. The dissociation of hydrated calcium oxalate is an endothermic process with an asymptotic limit of +470.84 kcal mol?1.
Figure
CaC2O4(H2O)16 and C2O4 2-(H2O)14  相似文献   

11.
The structures of ideal armchair (5,5) single-wall carbon nanotubes (SWCNTs) of different lengths (3.7, 8.8, and 16.0 Å for C40H20, C80H20, and C140H20) and with 1–10 hydroxyl groups at the end of the nanotube were fully optimized at the B3LYP/3-21G level, and in some cases at the B3LYP/6-31G* level, and the energy associated with the attachment of the OH substituent was determined. The OH-group attachment energy was compared with the OH functionalization of phenanthrene and picene models and with previous results for zigzag (9.0) SWCNT systems. In comparison to zigzag SWCNTs, the armchair form is more (by about 5 to 10 kcal mol?1) reactive toward hydroxylation.
Figure The structures of ideal armchair (5,5) single-wall carbon nanotubes (SWCNTs) of different lengths (3.7, 8.8, and 16.0 Å for C40H20, C80H20, and C140H20) and with 1–10 hydroxyl groups at the end of the nanotube were fully optimized at the B3LYP/3-21 G level, and in some cases at the B3LYP/6-31 G* level, and the energy associated with the attachment of the OH substituent was determined.
  相似文献   

12.
13.
Fully relativistic pseudo-potential ab-initio calculations have been performed to investigate the high pressure phase transition, elastic and electronic properties of lead-chalcogenides including the less known lead polonium. The calculated ground state parameters, for the rock-salt structure show good agreement with the experimental data. PbS, PbSe, PbTe and PbPo undergo a first-order phase transition from rock-salt to CsCl structure at 19.4, 15.5, 11.5 and 7.3 GPa, respectively. The elastic properties have also been calculated. The calculations successfully predicted the location of the band gap at L-point of Brillouin zone and the band gap for each material at ambient pressure. It is observed that unlike other lead-chalcogenides, PbPo is semi-metal at ambient pressure. The pressure variation of the energy gap indicates that these materials metalize under pressure. The electronic structures of these materials have been computed in parent as well as in high pressure B2 phase.
Figure
Equation of state of PbPo  相似文献   

14.
In this work, computations of density functional theory (DFT) were carried out to investigate the nature of interactions in solid 2,6-dibromo-4-nitroaniline (DBNA). This system was selected to mimic the hydrogen/halogen bonding found within crystal structures as well as within biological molecules. DFT (M06-2X/6-311++G**) calculations indicated that the binding energies for different of interactions lie in the range between ?1.66 and ?9.77 kcal mol?1. The quantum theory of atoms in molecules (QTAIM) was applied to provide more insight into the nature of these interactions. Symmetry-adapted perturbation theory (SAPT) analysis indicated that stability of the Br···Br halogen bonds is predicted to be attributable mainly to dispersion, while electrostatic forces, which have been widely believed to be responsible for these types of interactions, play a smaller role. Our results indicate that, for those nuclei participating in hydrogen/halogen bonding interactions, nuclear quadrupole resonance parameters exhibit considerable changes on going from the isolated molecule model to crystalline DBNA.
Figure
Electrostatic potential mapped on the surface of 2,6-dibromo-4-nitroaniline (DBNA) molecular electron density (0.001 e au?3). Color ranges for V S(r), in kcal?mol?1: red > 26.5, yellow 26.5–5.7, green 5.7– ?15.1, blue < ?15.1. Black circles Surface maxima, blue surface minima  相似文献   

15.
The present study reported the growth and metabolism characteristics of anaerobic ammonium-oxidizing (anammox) bacteria aggregates in an expanded granular sludge bed (EGSB). The results showed that the anammox bacteria aggregates presented starvation, growth, and inhibition phase along with the increase of substrate supply. The substrate conversion rates for survival were 0.05 kgNH 4 + –N/(kgVSS·day), 0.07 kgNO 2 ? –N/(kgVSS·day), and 0.12 kgN/(kgVSS·day); the substrate conversion rates for maximum growth were 0.21 kgNH 4 + –N/(kgVSS·day), 0.24 kgNH 4 + –N/(kgVSS·day), and 0.45 kgNH 4 + –N/(kgVSS·day), respectively. In the growth phase, the yield of anammox bacteria aggregates was 0.14 gVSS/(gNH 4 + –N), 0.12 gVSS/(gNO 2 ? –N), and 0.70 gVSS/(gNO 3 ? –N); the yield of extracellular polymeric substances (EPS) was 0.11 gEPS/(gNH 4 + –N), 0.09 gEPS/(gNO 2 ? –N), and 0.55 gEPS/(gNO 3 ? –N), respectively. The EPS contents in anammox bacteria aggregates were high compared to that in anaerobic granular sludge. Speculated from the cell yield, the energy for anammox bacteria growth was not only from nitrite oxidation, but also from anammox reaction.  相似文献   

16.
O-H…X and O-H…O H-bonds as well as C-X…X dihalogen and C-X…O halogen bonds have been investigated in halomethanol dimers (bromomethanol dimer, iodomethanol dimer, difluorobromomethanol…bromomethanol complex and difluoroiodomethanol…iodomethanol complex). Structures of all complexes were optimized at the counterpoise-corrected MP2/cc-pVTZ level and single-point energies were calculated at the CCSD(T)/aug-cc-pVTZ level. Energy decomposition for the bromomethanol dimer complex was performed using the DFT-SAPT method based on the aug-cc-pVTZ basis set. OH…O and OH…X H-bonds are systematically the strongest in all complexes investigated, with the former being the strongest bond. Halogen and dihalogen bonds, being of comparable strength, are weaker than both H-bonds but are still significant. The strongest bonds were found in the difluoroiodomethanol…iodomethanol complex, where the O-H…O H-bond exceeds 7 kcal mol-1, and the halogen and dihalogen bonds exceed 2.5 and 2.3 kcal mol-1, respectively. Electrostatic energy is dominant for H-bonded structures, in halogen bonded structures electrostatic and dispersion energies are comparable, and, finally, for dihalogen structures the dispersion energy is clearly dominant.
Figure
Competition of hydrogen, halogen, and dihalogen bonding in the bromomethanol dimer are investigated  相似文献   

17.
The crowned coumarin complexes are well known compounds for their ion recognition abilities. They undergo photophysical changes upon cation binding. On the basis of density functional theory calculations, we examined the sodium cation (Na+) binding energies of coumarin-crown ethers based on 15-Crown-5 (15 C5) and 18-Crown-6 (18 C6) as well as the optical absorptions of coumarin-crown ethers based on 12-Crown-4 (12 C4), 15 C5 and 18 C6. We explored why the attachment of crown ether ring to coumarin affects the Na+ binding energies of coumarin-crown ethers and also why the optical absorption of coumarin is modified by the crown ethers. Our study reveals that the Na+ ion binding energies of coumarin-crown ethers depend strongly on the size of the crown ether ring and also on the attachment position of the ether ring on coumarin. These factors affect the intramolecular charge transfer and overall stability of the complexes. The absorptions of the coumarin and ether ring parts of coumarin-crown ether are red shifted from those of isolated coumarin and crown ether, respectively. The red-shift of the coumarin ester group absorption is much stronger depending on the attachment position of the ether ring to coumarin. The absorption intensity of the coumarin part in coumarin-crown ethers is reduced for the benzene group absorption, but is enhanced for the ester group absorption.
Figure
In this study, the several Na+ - coumarin-crown ether coordination complexes based on 12-Crown-4, 15-Crown-5 and 18-Crown-6 were examined by the density functional theory method. The structural effects on the Na+ binding energies and optical properties were studied.  相似文献   

18.
DFT calculations for the acid-catalyzed hydrolysis of several maleamic acid amide derivatives revealed that the reaction rate-limiting step is determined on the nature of the amine leaving group. Further, it was established that when the amine leaving group was a secondary amine, acyclovir or cefuroxime moiety the tetrahedral intermediate formation was the rate-limiting step such as in the cases of acyclovir ProD 1- ProD 4 and cefuroxime ProD 1- ProD 4. In addition, the linear correlation between the calculated and experimental rates provided a credible basis for designing prodrugs for masking bitter taste of the corresponding parental drugs which have the potential to release the parent drug in a sustained release fashion. For example, based on the DFT calculated rates the predicted t1/2 (a time needed for 50 % of the reactant to be hydrolyzed to products) for cefuroxime prodrugs, cefuroxime ProD 1- ProD 4, were 12 min, 18 min, 200 min and 123 min, respectively.
Figure
A representation Scheme showing the interconversion of cefuroxime prodrug to cefuroxime by a prodrug chemical approach  相似文献   

19.
Mechanical and thermal properties of polyvinyl alcohol (PVA) are characterized and analyzed using in situ X-ray photoelectron spectroscopy (XPS) and quantum chemistry calculations. It is found that the carbon peaks—commonly used as the reference for spectroscopic analysis—shift under mechanical and thermal stretching. Results also indicate that, at different temperatures and among the various functional groups present in PVA, the carbon in the C–O group is the most stable. Computational calculations showed that Hartree–Fock/10-31G (d) reproduces the binding energy of core carbon electrons with an accuracy of 95 %, which is enough to characterize bonds, allowing the results of the spectroscopic analysis to be corroborated.
Figure
Complementary analysis of PVA-Ta-Lead film using X-ray photoelectron microscopy and ab initio quantum chemistry  相似文献   

20.
Density functional theory (DFT) calculations at B3LYP/6-31 G (d,p) and B3LYP/6-311?+?G(d,p) levels for the substituted pyridine-catalyzed isomerization of monomethyl maleate revealed that isomerization proceeds via four steps, with the rate-limiting step being proton transfer from the substituted pyridinium ion to the C=C double bond in INT1. In addition, it was found that the isomerization rate (maleate to fumarate) is solvent dependent. Polar solvents, such as water, tend to accelerate the isomerization rate, whereas apolar solvents, such as chloroform, act to slow down the reaction. A linear correlation was obtained between the isomerization activation energy and the dielectric constant of the solvent. Furthermore, linearity was achieved when the activation energy was plotted against the pK a value of the catalyst. Substituted-pyridine derivatives with high pK a values were able to catalyze isomerization more efficiently than those with low pK a values. The calculated relative rates for prodrugs 16 were: 1 (406.7), 2 (7.6?×?106), 3 (1.0), 4 (20.7), 5 (13.5) and 6 (2.2?×?103). This result indicates that isomerizations of prodrugs 1 and 35 are expected to be slow and that of prodrugs 2 and 6 are expected to be relatively fast. Hence, prodrugs 2 and 35 have the potential to be utilized as prodrugs for the slow release of monomethylfumarate in the treatment of psoriasis and multiple sclerosis.
Figure
Substituted pyridine-catalyzed isomerization of monomethylmaleate (prodrug, cis-isomer) to monomethylfumerate (parental drug, trans-isomer)  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