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 468 毫秒
1.
A new acentric mixed metal borate of composition K2Ba[B4O5(OH)4]2·10H2O, has been successfully obtained by slow evaporation solution method. The compound crystallizes in the Orthorhombic space group Pna21 (No.33) with a = 16.8668(7) Å, b = 13.0903(5) Å, c = 11.5529(5) Å and Z = 4. [B4O5(OH)4]2? clusters serve as fundamental building unit linking with BaO8, K1O6, K2O7 by common O atoms to form three-dimensional layer structure. The second harmonic generation measurements in the powder samples reveal that the compound exhibits approximately 0.5 times that of KH2PO4 (KDP) and phase matching.  相似文献   

2.
The infrared and Raman spectra, heat of formation (HOF) and thermodynamic properties were investigated by B3LYP/6-31G** method for a new designed polynitro cage compound 1,3,5,7,9,11-hexo(N(CH3)NO2)-2,4,6,8,10,12-hexaazatetracyclo[5,5,0,0,0]dodecane. The detonation velocity (D) and pressure (P) were predicted by the Kamlet–Jacobs equations based on the theoretical density and condensed HOF. The bond dissociation energies and bond orders for the weakest bonds were analysed to investigate the thermal stability of the title compound. The computational result shows that the detonation velocity and pressure of the title compound are superior to those of hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX), but inferior to those of 1,3,5,7-tetranitro-1,3,5,7-tetraazacyclooctane (HMX) and hexanitrohexaazaisowurtzitane (HNIW). And the analysis of thermal stability shows that the first step of pyrolysis is the rupture of the N7–NO2 bond. The crystal structure obtained by molecular mechanics belongs to the P21 space group, with the lattice parameters Z = 2, a = 11.8246 Å, b = 10.4632 Å, c = 15.9713 Å, ρ = 1.98 g cm?3.  相似文献   

3.
Molecular modeling of the cholesteric liquid crystal polyester poly[oxy(1,2 - dodecane)oxycarbonyl-1,4-phenyleneoxycarbonyl-1,4-phenylenecarbonyloxy-1,4-phenylenecarbonyl] (PTOBDME), [C34H36O8] n , synthesized in our laboratory and thermally characterized by differential scanning calorimetry (DSC), was performed to explain both its cholesteric mesophase and 3D crystalline structure. Conformational analysis (CA) was run for the monomer both by systematic search and with molecular dynamics (MD) simulations. Minima energy conformers were “polymerized” with Cerius2 and helical, cholesteric molecules were obtained in all cases. Our models agree with the chiral behavior observed by X-ray diffraction (XRD), thermooptical analysis (TOA) and circular dichroism (CD) experiments. Crystal packing of the polymer molecules were simulated in cells with parameters a and b obtained from experimental powder X-ray diffraction patterns and c calculated from the translational repetitive unit during the theoretical polymerization. Recalculated X-ray powder diffraction patterns of our models matched the observed ones. Morphology simulation from those crystal models is in good agreement with the crystals observed by optical microscopy. We have also modeled the self-associating nature of those polyesters when dispersed in aqueous media. Simulation of our models surrounded by different solvents, such as water and chloroform, were performed by calculating their interaction energies, coordination numbers, and mixing energies, applying Monte Carlo simulation techniques based on the Flory-Huggins theory. These results were compared with their experimental vibrational Fourier transform (FT)–Raman spectra in the regions in which structural marker bands of the polymer appear.  相似文献   

4.
The growth rate of thin polymer films polymerized in the positive column of a dc-glow discharge on the conditions [total pressure: Pneon + Pbenzene = 1 Torr, admixture of benzene : 10?4 < x < 10?1, discharge current: 1 mA — 15 mA] is calculated and measured. The calculation is based upon a previously developed polymerization model [6], which ascribe the major part of the growth rate to the insertion and crosslink of benzene ions into the substrate surface. The calculated growth rates of the polymer films and the separated portions of the considered polymerization processes correspond to the experiments within the bounds of the actual possibilities of analyzing the model.  相似文献   

5.
J.P. Colpa 《Molecular physics》2013,111(2):581-585
Measurements have been made of the neutron scattering structure factors of liquid nitrogen and liquid oxygen at 77 K and 84 K respectively in the Q-value range of 3 to 36 Å-1. ‘White’ incident thermal neutrons were produced in the wavelength range of 0·3 to 3·0 Å by a pulsed electron linac and detected in a total-scattering time-of-flight spectrometer. Qualitative agreement has been obtained between these present data and a simple molecular form factor in the Q-value range of 12–36 Å-1.

Using reactor data [1], structure factors over a total Q-value range of 0·3 to 36 Å-1 were obtained. Fourier transforms of the structure factor curves yield pair distribution functions which show a distinct separation of the inter and intra-molecular distances in the liquid. A comparison is made with results for the solid phase.  相似文献   

6.
Optical‐trapping confocal Raman microscopy allows the 1, 4‐addition reaction of diacetylenic functional groups in 1,2‐bis(10,12‐tricosadiynoyl)‐sn‐glycero‐3‐phosphocholine lipids to be monitored in individual phospholipid vesicles. Optical trapping allows a single vesicle to be observed over time, allowing the direct observation of structural changes in the vesicle membrane during polymerization. Confocal Raman microscopy excludes light collection outside the optical‐trap region avoiding interferences from the surrounding solution, while chemical reactions occurring in the membrane of the trapped vesicle can be measured with high sensitivity. Individual, optically trapped liposomes (0.6 µm in diameter) were exposed to photolysis radiation at 254 nm. Upon exposure to UV light, the cross‐linking polymerization reaction formed a conjugated ene–yne backbone in the bilayer of the optically trapped vesicle. Polymerization produces two different polymers, red and yellow in color, which can be distinguished structurally by their Raman spectra. Rates of red and yellow polymer formation were monitored by the Raman scattering intensities from both C = C stretching vibrations at 1455 cm–1 and 1508 cm–1 and C ≡ C stretching vibrations at 2080 and 2110 cm–1, respectively. Polymer formation rates depended linearly on 254‐nm light intensity, consistent with a one‐photon excited polymerization reacting in a photostationary state. Relative populations of red and yellow polymer in a polymerized vesicle depend sensitively on the sample temperature. From temperature‐dependent Raman spectra, the enthalpy change of the red‐to‐yellow thermochromic response and corresponding structural changes in the polymer could be determined. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

7.
A new inorganic-organic hybrid material produced from 2,6-dimethylanilinium cations and tin halide (SnCl6)2? has been synthesized and structurally determined by X-ray diffraction method. The title compound crystallizes in the monoclinic system, space group C2/m with a = 19.8772(4), b = 6.9879(1), c = 8.3001(2) Å, β = 98.487(2)° and V = 1140.26(4) Å3. The crystal structure is built up of sheets of (SnCl6)2? octahedral anions and 2,6-xylidinium cations. The optical band gap was calculated and found to be 4.11 eV. At high temperature this compound exhibits a structural phase transition at 338 K. This has been characterized by differential scanning calorimetric and dielectric studies. Measurements of AC conductivity as a function of frequency at different temperatures indicated the hopping conduction mechanism. The bioassay results showed that the structure exhibits significant antibacterial activity.  相似文献   

8.
In this work, a novel compound Bis(2-chloropropyl-N,N-dimethyl-1-ammonium) hexachloridostannate(IV) was synthesized and characterized by; single X-ray diffraction, Hirshfeld surface analysis, differential scanning calorimetric and dielectric measurement. The crystal structure refinement at room temperature reveled that this later belongs to the monoclinic compound with P21/n space group with the following unit cell parameters a = 7.2894(7) Å, b = 12.9351(12) Å, c = 12.2302(13) Å and β = 93.423 (6) °. The structure consists of isolated (SnCl6)2? octahedral anions connected together into layers via hydrogen bonds N–H….Cl between the chlorine atoms of the anions and the hydrogen atoms of the NH groups of the [C5H13NCl]+ cations. Hirschfeld surface analysis has been performed to gain insight into the behavior of these interactions. The differential scanning calorimetry spectrum discloses phase transitions at 367 and 376.7 K. The electrical properties of this compound have been measured in the temperature range 300–420 K and the frequency range 209 Hz–5 MHz. The Cole–Cole (Z′ versus Z″) plots are well fitted to an equivalent circuit model. The transition phase observed in the calorimetric study is confirmed by the change as function of temperature of electrical parameter such as the conductivity of grain (σg) and the σdc.  相似文献   

9.
Using high-purity starting materials, we synthesized a new room-temperature Cs2KInCl6 phase (I): monoclinic (C2/c), a = 25.484(11), b = 7.699(2) and c = 13.225(3) Å, β = 100.69(3)°. A ferroelasticparaelastic phase transition is noted at Tc = 100°C (by Thermal Analysis and X-ray, Raman-scattering, electrical permittivity versus temperature) leading to the prototype Fm3m phase(II) with a = 10.870(5) Å, quenchable when very slightly spoiled by impurities.  相似文献   

10.
To probe the molecular packing in crystalline domains of an unsymmetrical poly(benzoxazole-imide) (BPDA-BOA) introduced alternating phenylbenzoxazole and bisphthalimide units via the two-step polymerization of 5-amino-2-(4- aminophenyl)benzoxazole (BOA) and 3,3′,4,4′-biphenyltetracarboxylic dianhydride (BPDA), the crystal structure of its model compound, 5-phthalimido-2-(4-phthalimidophenyl)benzoxazole (PA-BOA), was investigated using powder wide-angle X-ray diffraction (WAXD) combined with molecular modeling using a Materials Studio program. Powder WAXD pattern of the model compound can be well indexed in terms of amonoclinic unit cell with parameters of a = 12.005 Å, b = 3.837 Å, c = 23.562 Å, β = 96.711°. There are two molecules in the unit cell with the space group of P2 (3). Based on the crystal structure of the model compound powders and the WAXD analysis of poly(benzoxazole-imide) film, it can be concluded that the projection of the monomer repeat length of the poly(benzoxazole-imide) chains in the chain directionis around 19.0 Å, and the interchain side-by-side distance and face-to-face distance between the centroids of two neighboring BPDA-BOA chains are around 6.10 and 4.01 Å, respectively. The difference of molecular packing between the poly(benzoxazole-imide) chains in their crystalline domains and the model compounds in their crystals are also presented. Both the side-by-side and face-to-face π-π stacking distances between two neighboring polymer chains in the poly(benzoxazole-imide) crystalline domains are significantly larger than the corresponding values between two neighboring molecules in the model compound crystals due to a more kinked and twisted conformation.  相似文献   

11.
Inspired by the works of Rodnianski and Schlein [31] and Wu [34,35], we derive a new nonlinear Schrödinger equation that describes a second-order correction to the usual tensor product (mean-field) approximation for the Hamiltonian evolution of a many-particle system in Bose-Einstein condensation. We show that our new equation, if it has solutions with appropriate smoothness and decay properties, implies a new Fock space estimate. We also show that for an interaction potential ${v(x)= \epsilon \chi(x) |x|^{-1}}Inspired by the works of Rodnianski and Schlein [31] and Wu [34,35], we derive a new nonlinear Schr?dinger equation that describes a second-order correction to the usual tensor product (mean-field) approximation for the Hamiltonian evolution of a many-particle system in Bose-Einstein condensation. We show that our new equation, if it has solutions with appropriate smoothness and decay properties, implies a new Fock space estimate. We also show that for an interaction potential v(x) = ec(x) |x|-1{v(x)= \epsilon \chi(x) |x|^{-1}}, where e{\epsilon} is sufficiently small and c ? C0{\chi \in C_0^{\infty}} even, our program can be easily implemented locally in time. We leave global in time issues, more singular potentials and sophisticated estimates for a subsequent part (Part II) of this paper.  相似文献   

12.
Molecular structure, vibrational energy levels and potential energy distribution of 1H‐imidazo[4,5‐b]pyridine, 3H‐imidazo[4,5‐b]pyridine, 5‐methyl‐1H‐imidazo[4,5‐b]pyridine, 6‐methyl‐1H‐imidazo[4,5‐b]pyridine and 7‐methyl‐3H‐imidazo[4,5‐b]pyridine were determined using density functional theory (DFT) at the B3LYP/6‐31G(d,p) level. The optimised bond lengths and bond angles are in good agreement with the X‐ray data of 5‐methyl‐1H‐imidazo[4,5‐b]pyridine obtained in the present work (Pbca space group; a = 8.660(2), b = 11.078(2), c = 11.078(3) Å, Z = 8). The N+H group plays the role of a proton donor in a medium strong hydrogen bond of the type N H…N, linking the N‐atom of the pyridine with the adjacent molecule related by the symmetry operation: 1/2 − x, y − 1/2, z(N…N = 2.869(25) Å). The presence of hydrogen bond is confirmed by appearance in the IR spectra of a very broad and strong contour in the 2000–3100 cm−1 range. The place of substitution of the methyl group at the pyridine ring influences the proton position of the NH group at the imidazole unit. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

13.
We obtained a new complex containing sarcosine (CH3NH2+CH2COO) and barium(2 + ) dibromide (TSBB) in 3:1 molar ratio, as well as its deuterated analog. Single‐crystal X‐ray diffraction measurements show that TSBB crystallizes in the monoclinic system, space group P2(1)/c. The unit cell parameters are as follows: a = 18.345(4) Å, b = 10.668(2) Å, c = 8.9212(18) Å, β = 91.86(3)°, and Z = 4. The structure was determined with final R1 = 0.0396 (for I > 2σI). The crystal possesses a pseudohexagonal symmetry down c axis showing the resemblance to the crystal structure of trissarcosine calcium chloride (TSCC). There are N HBr hydrogen bonds (HB) of six types. TSBB crystal undergoes a phase transition at 416 K (heating)–415 K (cooling) of continuous nature. The spectroscopic [Infrared (IR) and Raman] investigation of the crystal was performed at room temperature. The results are discussed with respect to the crystallographic data, as well as the results obtained for TSCC crystal. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

14.
The crystal structure of the paraelectric phase of rubidium hydrogen sulfate has been redetermined at room temperature to be monoclinic with a = 14.3503(14), b = 4.6187(4), c = 14.3933(14)?Å, β = 118.03(1)° (space group P21/n). Both the sulfate groups are found to be ordered, unlike in previous reports. The crystal structure of the ferroelectric phase at 200?K belongs to the noncentrosymmetric space group Pn with a = 14.2667(12), b = 4.5878(4), c = 14.2924(12)?Å, β = 118.01(1)°, with distorted sulfate groups. The change in the Rb coordination is discussed in terms of bond-valence-sum calculations. Variable-temperature powder X-ray diffraction patterns at temperatures above 393?K indicate a possible reduction in symmetry, suggesting a phase transition.  相似文献   

15.
A series of Zn(II)-Schiff bases I, II and III complexes were synthesized by reaction of o-phenylenediamine with 3-methylsalicylaldehyde, 4-methylsalicylaldehyde and 5-methylsalicylaldehyde. These complexes were characterized using FT-IR, UV-Vis, Diffuse reflectance UV-Vis, elemental analysis and conductivity. Complex III was characterized by XRD single crystal, which crystallizes in the triclinic system, space group P-1, with lattice parameters a?=?9.5444(2) Å, b?=?11.9407(2) Å, c?=?21.1732(3) Å, V?=?2390.24(7) Å3, D c ?=?1.408 Mg m?3, Z?=?4, F(000)?=?1050, GOF?=?0.981, R1?=?0.0502, wR2?=?0.1205. Luminescence property of these complexes was investigated in DMF solution and in the solid state. Computational study of the electronic properties of complex III showed good agreement with the experimental data.  相似文献   

16.
Abstract

Vinyl benzoate was polymerized by γ-radiation from a cobalt-60 source and the kinetics of polymerization were studied at several temperatures. The results showed that the rate of polymerization was proportional to I0.66 , where I is the radiation dose rate. The net activation energy for the polymerization reaction, (Ep -1/2 Et ), was found to be 3.62 kcal, where Ep and Et are the activation energies for the propagation and termination stages of the reaction. The radical yield measured by the loss of DPPH in the solution after irradiation was G=5.0, while the G (Radical) effective in initiating polymerization was 0.94. The ratio of the rate constants, k2 p/kt = 5.8 × 10?4 at 60° and 1.59 × 10?4 at 25° The rate of polymerization was higher than that of styrene but lower than that of vinyl acetate under comparable conditions.  相似文献   

17.
First principles molecular orbital and plane‐wave ab initio calculations have been used to investigate the structural and energetic properties of a new cage compound 2, 4, 6, 8, 12‐pentanitro‐10‐(3, 5, 6‐trinitro (2‐pyridyl))‐2, 4, 6, 8, 12‐hexaazatetracyclo [5.5.0.03,11.05,9]dodecane (PNTNPHATCD) in both the gas and solid phases. The molecular orbital calculations using the density functional theory methods at the B3LYP/6‐31G(d,p) level indicate that both the heat of formation and strain energy of PNTNPHATCD are larger than those of 2, 4, 6, 8, 10, 12‐hexanitro‐2, 4, 6, 8, 10, 12‐hexaazatetracyclo [5.5.0.0.0] dodecane (CL‐20). The infrared spectra and the thermodynamic property in gas phase were predicted and discussed. The calculated detonation characteristics of PNTNPHATCD estimated using the Kamlet–Jacobs equation equally matched with those of CL‐20. Bond‐breaking results on the basis of natural bond orbital analysis imply that C–C bond in cage skeleton, C–N bond in pyridine, and N–NO2 bond in the side chain of cage may be the trigger bonds in the pyrolysis. The structural properties of PNTNPHATCD crystal have been studied by a plane‐wave density functional theory method in the framework of the generalized gradient approximation. The crystal packing predicted using the Condensed‐phase Optimized Molecular Potentials for Atomistic Simulation Studies (COMPASS) force fields belongs to the Pbca space group, with the lattice parameters a = 20.87 Å, b = 24.95 Å, c = 7.48 Å, and Z = 8, respectively. The results of the band gap and density of state suggest that the N–NO2 bond in PNTNPHATCD may be the initial breaking bond in the pyrolysis step. As the temperature increases, the heat capacity, enthalpy, and entropy of PNTNPHATCD crystal all increase, whereas the free energy decreases. Considering that the cage compound has the better detonation performances and stability, it may be a superior high energy density compound. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

18.
《光谱学快报》2013,46(6):553-564
Abstract

The cis stereochemistry of 6‐(4‐methoxy‐phenyl)‐1,5,7a‐triphenyl‐tetrahydro‐imidazo[1,5‐b][1,2,4]oxadiazol‐2‐one was studied by use of a PM3 semi‐empirical quantum mechanical model, and x‐ray crystallographic analysis. It crystallizes in the monoclinic space group P2 1 /n with a = 10.812(1) Å, b = 16.464(2) Å, c = 13.379(1) Å, α = 90.00°, β = 98.39(1)°, γ = 90.00°, V = 2356.07(4) Å3, Z = 4, D calc = 1.3067 g cm?3, F(0 0 0) = 976.41, and μ = 0.086 mm?1. The structure was solved by direct methods and refined to R = 0.066 for 1257 independent reflections [I > 4σ (I)]. The results from x‐ray diffraction were seen to be generally consistent with the results from previously reported spectroscopic investigations, beside theoretical calculations, except for conformations of five‐membered fused heterocycles. Two inter‐ and intramolecular weak interactions in addition to carbon atoms (C1 and C3) with different chiralities were found in the structure. The conformational study was performed by randomly scanning the potential energy surface belonging to the title compound with respect to selected torsion angles.  相似文献   

19.
Four novel coordination polymers: Ag(dpa) I, Co(O3PH)(4,4′-bpy)(H2O) II, Zn(O3PH)(4,4′-bpy)0.5 III and Mn[O2PH(C6H5)]2(4,4′-bpy) IV (dpa=2,2′-dipyridylamine; 4,4′-bpy=4,4′-bipyridine), were synthesized by microwave heating and characterized by X-ray crystallography. I crystallizes in monoclinic space group P21/n with a=11.576(2) Å, b=5.585(2) Å, c=15.243(4) Å, β=109.00(2)°, V=931.8(3) Å3. II crystallizes in monoclinic Cc space group with a=22.477(7) Å, b=5.280(1) Å, c=10.404(4) Å, β=96.08(3)°, V=1227.8(7) Å3. III crystallizes in monoclinic P21/c space group with a=9.758(2) Å, b=7.449(3) Å, c=10.277(2) Å, β=100.02(2)°, V=735.6(4) Å3. IV crystallizes in monoclinic space group P2/c with a=10.174(1) Å, b=11.817(3) Å, c=18.784(4) Å, β=102.14(1)°, V=2207.8(8) Å3. I consists of linear metal–metal chains wrapped by dpa ligands. II and III consist of two-dimensional MII(O3PH) inorganic sheets cross-linked by 4,4′-bpy ligands, while IV is formed by Mn[O2PH(C6H5)]2 sheets cross-linked by 4,4′-bpy ligands. I exhibits two-step thermal decomposition at ~200 and ~250°C, resulting in the reduction of Ag+ to Ag metal. II loses its coordination water at ~100°C, leaving vacant coordination sites at Co2+ ions, while the original framework remains intact. The removal of 4,4′-bpy in IIIV occurs at elevated temperatures above 250, 200 and 400°C respectively.  相似文献   

20.
ABSTRACT

Perovskite structured mixed metal fluorides containing manganese/sodium or potassium have been synthesized in pure form by a greener precipitation route and characterized by high-resolution powder X-ray diffraction and Raman spectroscopy techniques. While all the reflections in the powder X-ray diffraction pattern of potassium manganese fluoride could be indexed in cubic symmetry with a = 4.1889 Å, sodium manganese fluoride showed reflections at positions typical of orthorhombic symmetry (Pnma space group) with a = 5.751, b = 8.008, and c = 5.548 Å. Potassium manganese fluoride in powder form showed bands at 209, 291, 386, 558, 621, and 733 cm?1 in the Raman spectrum at room temperature. All these bands disappeared and second-order band at 1151 and 1298 cm?1 emerged when the powders were compacted under pressure ranging between 1 and 4 tons (uniaxial). A similar change was noticed for sodium manganese fluoride in which bands at 1099, 1149, 1203, and 1286 cm?1 were observed for the compacted samples. The response of the vibrational modes of these compounds to uniaxial pressure revealed the existence of large structural disorder in them. Additionally, the need for the extreme care to collect and interpret Raman data of polycrystalline samples of these systems has been illustrated through this study.  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