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.
The Becke3LYP functional of DFT theory and the two-layered ONIOM (B3LYP/6–311 + G(d,p): MNDO) method were used to characterize 46 gas-phase complexes of 34 neutral and anionic ligands (H2O, CH3OH, CH3COOH, CH3CONH2, HOSO2NH2, CO2, HSO2NH2, CH3SO2NH2, CH3C(= O)NHOH, imidazole, NH2SO2NH2, anions of 4-aminobenzenesulphonamide, saccharin, 1I9L, brinzolamide, dorzolamide, acetazolamide, further HO(?), CH3CO(?), CH3COO(?), CH3CONH(?), N = N = N(?), S = C = N(?), CH3C(= O)NHO(?), HOCOO(?), imidazoleN(?), phenol-O(?), HOSO2NH(?), (?)OSO2NH(?), (?)OSO2NH2, H2NSO2NH(?), HSO2NH(?), CH3SO2NH(?), and CF3SO2NH(?), respectively) with Zn2+. Proton dissociation enthalpies and Gibbs energies of acidic inhibitors in the presence of zinc were computed. Their gas-phase acidity considerably increases upon chelation. Of the bases investigated, the weakest zinc affinity is exhibited by carbon dioxide (- 313.5 kJ mol?1). Deprotonated inhibitors have higher affinities for zinc than the neutral ones. Compared to the other mono-deprotonated ligands the acetohydroxamic acid anion has the highest affinity for zinc (- 1872.7 kJ mol?1). The zinc affinity of the acetazolamide anion computed using the hybrid ONIOM (B3LYP/6-311 + G(d,p): MNDO) method is in very good agreement with the full DFT ones and this method can be adopted to model large complexes of inhibitors with the active site of carbonic anhydrase.  相似文献   

2.
Microwave spectra of NCCCH–NH3, CH3CCH–NH3, and NCCCH–OH2have been recorded using a pulsed-nozzle Fourier-transform microwave spectrometer. The complexes NCCCH–NH3and CH3CCH–NH3are found to have symmetric-top structures with the acetylenic proton hydrogen bonded to the nitrogen of the NH3. The data for CH3CCH–NH3are further consistent with free or nearly free internal rotation of the methyl top against the ammonia top. For NCCCH–OH2, the acetylenic proton is hydrogen bonded to the oxygen of the water. The complex has a dynamicalC2vstructure, as evidenced by the presence of two nuclear-spin modifications of the complex. The hydrogen bond lengths and hydrogen-bond stretching force constants are 2.212 Å and 10.8 N/m, 2.322 Å and 6.0 N/m, and 2.125 Å and 9.6 N/m for NCCCH–NH3, CH3CCH–NH3, and NCCCH–OH2, respectively. For the cyanoacetylene complexes, these bond lengths and force constants lie between the values for the related hydrogen cyanide and acetylene complexes of NH3and H2O. The NH3bending and weak-bond stretching force constants for CH3CCH–NH3are less than those found in NCCCH–NH3, NCH–NH3, and HCCH–NH3, suggesting that the hydrogen bonding interaction is particularly weak in CH3CCH–NH3. The weakness of this hydrogen bond is partially a consequence of the orientation of the monomer electric dipole moments in the complex. In CH3CCH–NH3the antialigned monomer dipole moments lead to a repulsive dipole–dipole interaction energy, while in NCH–NH3and NCCCH–NH3the aligned dipoles give an attraction interaction.  相似文献   

3.
The room-temperature elastic constants of a number of hexahalometallate A2MX6 single crystals [K2SnCl6, K2ReCl6, (NH4)2SnCl6, (NH4)2SnBr6, (NH4)2SiF6, Rb2SnBr6, K2SeBr6, (NH4)2TeBr6, K2PtBr6 and (NH4)2PtBr6] have been measured either by Brillouin scattering or by the ultrasonic pulse echo overlap technique. Refractive indices have also been determined. These antifluorite structure compounds contain large MX2?6 ions and the interionic spacings are much greater than those of the alkaline-earth fluorite structure halides: their elastic stiffnesses are correspondingly smaller. Hydrostatic pressure derivatives of the elastic stiffness constants have been measured for K2SnCl6, (NH4)2SnBr6 and (NH4)2SnCl6 and are found to be positive; there is no marked softening of the long-wavelength acoustic-phonon modes at room temperature. The vibrational anharmonicities of these long-wavelength modes are discussed in terms of the acousticmode Grüneisen parameters, which are compared with the thermal Grüneisen parameters. For K2SnCl6 a mean of optic- and acoustic-mode Grüneisen parameters is shown to correlate well with the thermal Grüneisen parameter.  相似文献   

4.
A low-temperature all-solid-state thermal method has been developed to synthesize electrolytes like LiAsF6, LiPF6 and allied lithium-based fluoro-chemicals useful for lithium secondary cells. This developed procedure is simple and environmentally friendly compared with the existing procedures, which are poisonous and hazardous to health due to the use of obnoxious gases or liquids like F2, AsF3, PF3, BF3, AsF3, or AsF5, which are difficult to handle. In this proposed procedure, all chemicals are taken as analytical grade. Lithium salts and all other required salts are mixed well in required proportions and heated in between 300–400 °C to get the products examined by X-ray. The chemical reactions proposed for the preparation are given below.
  1. 2LiOH + As2O5 + 12NH4F → 2LiAsF6 + 12NH3 + 7H2O
  2. LiOH + (NH4)2HPO4 + 6NH4F → LiPF6 + 8NH3 + 5H2O
  3. 2LiOH + B2O3 + 8NH4F → 2LiBF4 + 8NH3 + 5H2O
  相似文献   

5.
《Infrared physics》1992,33(6):589-592
The vibrational anharmonicity and Grüneisen parameters of hexahelometallate A2MX6 single crystals have been determined theoretically by making use of phonon lattice theory. The potential model employed to calculate these properties consists of long range coulomb, three body interactions, short range overlap repulsion effective upto the nearest neighbour ions and phonon-lattice interactions. These antifluorite structure compounds contain large MX2−6- ions and as the interionic spacings are much greater than those of the alkaline-earth fluorite structure halides, their elastic constants are correspondingly smaller. The hydrostatic pressure derivatives of the second order elastic constants (SOEC) calculated for K2SnCl6, K2ReCl6, (NH4)2SnCl6, (NH4)2TeCl6, (NH4)2SnBr6, and (NH4)2TeBr6, are found to be positive and close to the experimental values. The vibrational anharmonicities of the long-wavelength modes are explained in terms of the acoustic mode Grüneisen parameters.  相似文献   

6.
《Surface science》1986,167(1):207-230
A unified electron spectroscopic study of polycrystalline Ti and its interaction with H2, O2, N2, and NH3 are described. Auger electron spectroscopy (AES), electron energy-loss spectroscopy (ELS), ultraviolet and X-ray photoelectron spectroscopy (UPS and XPS) are combined to provide detailed information about the electronic structure of the titanium surface and its interaction with these adsorbates. X-ray and ultraviolet photoelectron spectra and electron energy-loss spectra are presented for the clean titanium surface, and following exposure to H2, O2, N2 and NH3. Spectral assignments are provided in each case. The electron spectra of oxygen exposed Ti and nitrogen sputtered Ti are quite similar, and are interpreted with reference to band structure calculations for TiO and TiN. Electron spectroscopy indicates essentially complete dissociative adsorption of NH3 on the clean titanium surface.  相似文献   

7.
This paper reports that the equilibrium structure of NH2 has been optimized at the QCISD/6-311++G (3df, 3pd) level. The ground-state NH2 has a bent (C2v, X^2B1) structure with an angle of 103.0582°. The geometrical structure is in good agreement with the other calculational and experimental results. The harmonic frequencies and the force constants have also been calculated. Based on the group theory and the principle of microscopic reversibility, the dissociation limits of NH2(C2v, X^2B1) have been derived. The potential energy surface of NH2(X^2B1) is reasonable. The contour lines are constructed, the structure and energy of NH2 reappear on the potential energy surface.  相似文献   

8.
TH Freeda  C Mahadevan 《Pramana》2001,57(4):829-836
Pure and impurity added (with NH4Cl, NH4NO3, NH4H2PO4, and (NH4)2SO4) KDP single crystals were grown by the gel method using silica gels. X-ray diffraction data were collected for powder samples and used for the estimation of lattice variation and thermal parameters like Debye-Waller factor, mean-square amplitude of vibration, Debye temperature and Debye frequency. The thermal parameters do not vary in a particular order with respect to impurity concentration. The results obtained are reported and discussed.  相似文献   

9.
Usingab initio data of the potential surface for the ground state of NH3 a function is obtained to represent, with a very high precision, the potential along the inversion coordinate. The inversion spectra of NH3, ND3, NH2D and NHD2 are found by numerically solving the Schrödinger equation for this potential. Comparison is made of the calculated inversion frequencies and the experimental values, and the molecular constants of NH3 are also compared to those found in the literature.  相似文献   

10.
The magnetic susceptibility of the metallic compounds Ca(NH3)6, Sr(NH3)6 and Ba(NH3)6 has been measured by the Faraday method in the range 2–200 K. The susceptibility of Ca(NH3)6 is strongly dependent on temperature and exhibits a broad minimum near 120 K and a peak near 10 K. In contrast to Ca(NH3)6, the susceptibilities of Sr(NH3)6 and Ba(NH3)6 are diamagnetic and decrease rapidly as the temperature is lowered. At low temperatures Ba(NH3)6 has the largest mass susceptibility of any nonsuperconducting metal.  相似文献   

11.
The FT-Raman spectrum of cupferron, [PhN2O2]NH4 and the micro-Raman spectra of the new corresponding cobalt(II) cupferronato complexes, CoL2A2, L = PhN2O2, a = H2O, MeOH, o-C6H4(NH2)2, p-C6H4(NH2)2 and CoL2A, a = (-C6H4NH2-p)2 were recorded and discussed. All the complexes show a Raman band at about 1302 cm1 and the characteristic v(N-N) and δ(ONNO) modes of the anionic ligand. the vibrational analysis of the title compounds reveals the electron delocalisation over the N-nitroso-N-hydroxylaminato (ONNO) unit, as well as the bidentate coordination of the cupferronato ligand to the metal center through the oxygen atoms.  相似文献   

12.
Gradual substitution of coal with green ammonia is a practical approach for the coal power phasedown at a minimal cost of modification, but the ignition and gas-phase reaction during co-firing NH3 with coal remain largely unclear. In this work, we investigate the co-combustion behaviors of NH3 and a high-volatile coal on a two-stage flat flame burner. Remarkably, the post-flame oxygen mole fraction Xi,O2 of the inner stage can be manipulated to reproduce a proper reducing-to-oxidizing environment that coal particles experience in the practical combustor. We first reveal that, under certain values of Xi,O2 and NH3 co-firing energy ratios ENH3, the reaction intensity (manifested by OH-PLIF signals) in the NH3-coal flame is stronger than burning either pure coal or NH3. This synergetic effect originates from an NH3-combustion-induced enhancement of volatile release. We then propose a characteristic time scale τOH from the OH signals for the initiation of overall reactions in the system. In the case of Xi,O2=0, τOH monotonically increases with ENH3, while for Xi,O2=0.2, the trend transitions to a decreasing one. It can be interpreted by comparing τOH with the characteristic O2 diffusion time, coal particle heating time, and the coal pyrolysis time under different Xi,O2. Furthermore, the coal particle ignition in coal-NH3 flames can no longer be determined by visual images. Instead, we apply CH* chemiluminescence to identify the stages of coal particle ignition and volatile combustion in the NH3-coal flame. While NH3 addition has both positive (elevating temperatures & diluting coal particles) and negative (consuming O2) effects on coal ignition, the combined influence of ENH3 is marginal on coal ignition delay time. On the other hand, the volatile combustion time decreases linearly with ENH3, suggesting a pure effect of reduced coal feed rate.  相似文献   

13.
The change of the discharge voltage when laser light crossing the discharge is tuned to a molecular transition has been measured. Experiments have been performed in the wavelength region between 570 nm and 620 nm with discharges in NH3, NO2, H2, N2, O2 and argon. Transitions from the ground states of NH2 and NO2 and transitions from metastable states of N2 and H2 have been detected. The spacial dependence of the opto galvanic in a low pressure dc-discharge of H2 and N2 has been studied.  相似文献   

14.
The infra-red spectra of KH2PO4, NH4H2PO4, NaH2PO4, KH2AsO4, NH2H2AsO4, Ag2H3IO6 and (NH4)2H3IO6 and of their deuterated analogues have been recorded at room temperature and some of them also at low temperature in the ferroelectric phase. The interpretation of the particularly interesting spectral region between 3000 and 1500 cm?1, containing several OH bands, has been made in terms of the tunnelling of the protons between two minima of potential energy. These were taken to be of equal depth in the non-ferroelectric phase and unsymmetrical in the ferroelectric form. A quantum-mechanical treatment of the vibrational problem of the latter type has been carried out and is presented in the Appendix. Good agreement has been found between the theoretically predicted energy levels and the experimental data.  相似文献   

15.
The He(I) photoelectron spectrum of dibromamine, NHBr2, has been obtained from observation of the products of the room temperature gas-phase reaction of NH3 and Br2. Due to its extreme reactivity, only a small proportion of NHBr2 is present in various mixtures containing NH2Br and N2 as well as unreacted NH3 and Br2. Careful spectrum stripping is required to obtain a spectrum of NHBr2 which is identified by the relationship of its spectrum to those of NH2Br, NH2Cl and NHCl2.  相似文献   

16.
Starting from force constant values calculated by an ab initio MO method (4-31G(N1)), and by adjusting the diagonal elements, a practical force constant matrix (F) has been reached which could explain the observed infrared and Raman spectra (in the frequency range lower than 2000 cm?1) of the gauche form of the ethylamine CH3CH2NH2 molecule and five isotopic species CH313CH2NH2, CH3CH215NH2, CH3CD2NH2, CH3CH2ND2, and CD3CD2NH2. The F matrix for the trans form of ethylamine was constructed by transferring ab initio 4-31G(N1) values and by revising diagonal elements with conversion factors whose values are equal to the corresponding values of gauche form. A nearly complete set of assignments was achieved of the vibrational bands of ethylamines, observed so far in the spectral range 2000–100 cm?1. In matrix isolation spectroscopy, two bands assignable to the NH2 wagging vibrations of gauche and trans forms have been found at 775 and 782 cm?1, respectively, for CH3CH2NH2. They are at 768 and 774 cm?1, respectively, for CD3CD2NH2. From the intensity changes of these bands observed on changing the nozzle temperature in the matrix formation, the energy difference ΔE (gauche-trans) of these two conformers has been estimated to be 100 ± 10 cm?1.  相似文献   

17.
Deep insights into the combustion kinetics of ammonia (NH3) can facilitate its application as a promising carbon-free fuel. Due to the low reactivity of NH3, experimental data of NH3 combustion can only be obtained within a limited range. In this work, nitrous oxide (N2O) and hydrogen (H2) were used as additives to investigate NH3 auto-ignition in a rapid compression machine (RCM). Ignition delay times for NH3, NH3/N2O blends, and NH3/H2 blends were measured at 30 bar, temperatures from 950 to 1437 K. The addition of N2O and H2 ranged from 0 to 50% and 0 to 25% of NH3 mole fraction, respectively. Time-resolved species profiles were recorded during the auto-ignition process using a fast sampling system combined with a gas chromatograph (GC). An NH3 combustion model was developed, in which the rate constants of key reactions were constrained by current experimental data. The addition of N2O affected the ignition of NH3 primarily through the decomposition of N2O (N2O (+M) = N2 + O (+M), R1) and direct reaction between N2O and NH2 (N2H2 + NO = NH2 + N2O, R2). The rate constant of R2 was constrained effectively by experimental data of NH3/N2O mixtures. Two-stage ignition behaviors were observed for NH3/H2 mixtures, and the corresponding first-stage ignition delay times were reported for the first time. Experimental species profiles suggested the first-stage ignition resulted from the consumption of H2. The oxidation of H2 provided extra HO2 radicals, which promoted the production of OH radicals and initiated first-stage ignition. Reactions between HO2 radicals and NH3/NH2 dominated the first-ignition delay times of NH3/H2 mixtures. Moreover, the first-stage ignition led to the fast production of NO2, which acted as a key intermediate and affected the following total ignition. Consequently, the reaction NH2 + NO2 = H2NO + NO (R3) was constrained by total ignition delay times.  相似文献   

18.
The adsorption properties of a variety of atomic species (H, O, N, S, and C), molecular species (N2, HCN, CO, NO, and NH3) and molecular fragments (CN, NH2, NH, CH3, CH2, CH, HNO, NOH, and OH) are calculated on the (111) facet of palladium using periodic self-consistent density functional theory (DFT–GGA) calculations at ¼ ML coverage. For each species, we determine the optimal binding geometry and corresponding binding energy. The vibrational frequencies of these adsorbed species are calculated and are found to be in good agreement with experimental values that have been reported in literature. From the binding energies, we calculate potential energy surfaces for the decomposition of NO, CO, N2, NH3, and CH4 on Pd(111), showing that only the decomposition of NO is thermochemically preferred to its molecular desorption.  相似文献   

19.
Modulated charge separation across (MO)/CH3NH3PbI3 and (MO)/PbI2/CH3NH3PbI3 (MO = TiO2, MoO3) interfaces was investigated by surface photovoltage (SPV) spectroscopy. Perovskite layers were deposited by solution‐based one‐step preparation and two‐step preparation methods. An unreacted PbI2 layer remained at the interface between the metal oxide and CH3NH3PbI3 for two‐step preparation. For the two‐step preparation on TiO2, the SPV signal related to absorption in CH3NH3PbI3 increased in comparison to the one‐step preparation due to electron transfer from CH3NH3PbI3 via PbI2 into TiO2 whereas the SPV signal related to defect transitions decreased. For the one‐step preparation on MoO3, holes photogenerated in CH3NH3PbI3 recombined with electrons in MoO3. In contrast, a hole transfer from CH3NH3PbI3 towards MoO3 was blocked by the PbI2 interlayer for the two‐step preparation on MoO3. (© 2014 WILEY‐VCH Verlag GmbH &Co. KGaA, Weinheim)  相似文献   

20.
《Applied Surface Science》1986,27(3):275-284
The dissociation rates of H2, C2H4, C2H4, and NH3 have been studied on oxygen covered Pd surfaces by measuring the water desorption rates during exposure to each of the molecules. These results are correlated with the hydrogen response of a Pd-MOS structure. The measurements show a trend (at 473 K) where oxygen blocks H2 dissociation, blocks C2H4 dissociation only above a certain oxygen coverage, has no influence on C2H2 dissociation, and promotes NH3 dissociation.  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