首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 187 毫秒
1.
2,7-二甲基-2,4,6-辛三烯-1,8-二醛1(2,7-dimethylocta-2,4,6-triene-1,8-dial)一直以来都被认为是合成类胡萝卜素的关键中间体之一,国外对它的合成研究起步较早,国内在2005年之前鲜有报道.  相似文献   

2.
采用分子动力学方法,在正侧(NVT)系综下研究了N-脒基脲二硝酰胺盐(FOX-12)在溶剂中的晶体形貌.通过构建溶剂分子层-晶面的界面吸附模型模拟其动力学平衡构型,计算溶剂与晶体表面间的结合能,进而对真空附着能进行修正并获得溶剂条件下的晶貌.使用自然冷却法在水和水/甲醇中培养FOX-12晶体并利用扫描电子显微镜进行了表征.结果表明,在真空条件下决定FOX-12晶貌的6个重要晶面为(110),(200),(201),(011),(002)和(111);FOX-12在水溶液条件下的主要晶面为(110)和(011),在水/甲醇溶液条件下的主要晶面为(200)和(011),预测的晶体形貌与实验结果相吻合.对水分子和FOX-12的(110)面间的径向分布函数进行了计算,分析了水分子和晶面间的分子间作用力.  相似文献   

3.
聂福德  徐蓉  范仲勇  李越生 《色谱》2007,25(1):11-15
采用反相气相色谱技术(IGC)研究了4种不同粒度的1,3,5-三氨基-2,4,6-三硝基苯(TATB)的表面性质。4种不同粒度的TATB表面自由能的色散分量(γds)随着温度的升高而增加;粒度越大的粒子,其色散自由能上升越快;在较高温度下,粗颗粒TATB显示了更强的色散作用(γds=193.2 mJ/m2,353 K),粒度最小的亚微米TATB显示了最弱的色散作用(γds=64.0 mJ/m2,353 K)。由于制备方法不同和粒子大小的差异,4种TATB的表面酸碱性质显示了明显的差别,细颗粒TATB表面有较强的亲电子特性;而其他3种TATB在极性探针分子的作用下的吸附均表现为吸热吸附,表现出在分子内和分子间具有强烈的相互作用,其Ka和Kb值均为负。  相似文献   

4.
采用COMPASS力场和NVT正则系综的动力学模拟方法, 搭建了聚合度分别为10, 50和100的偏氟乙烯(VDF)/三氟氯乙烯(CTFE)交替共聚物, 对交替共聚物在1,3,5-三氨基-2,4,6-三硝基苯(TATB)的(0,0,1)晶面上的吸附和结构进行了分子动力学(MD)模拟. 结果表明, 在300~320 K温区, 聚合度为100的VDF/CTFE交替共聚物链对TATB晶体有理想的表面活性和吸附能力, 以train型构象平铺于TATB表面. 通过对聚合度为10的交替共聚物的多链体系在TATB表面吸附的MD模拟, 表明了VDF/CTFE交替共聚物具有非凝聚吸附的高表面活性特征. 对搭建的乙酸乙酯溶剂化的聚合度为50的VDF/CTFE交替共聚物在TATB晶体表面吸附的模拟, 实验证明了溶剂小分子能够降低共聚物链的吸附能力, 且链以tail型构象吸附于TATB表面.  相似文献   

5.
本文通过Suzuki偶联反应高效合成了两种三萘基三嗪化合物,即2,4,6-三(1-萘基)-1,3,5-三嗪(T1NT)和2,4,6-三(2-萘基)-1,3,5-三嗪(T2NT). 考察了介质的极性、温度以及THF-H2O二元溶剂体系中的溶解性等因素对它们吸收和发射光谱行为的影响.研究发现, 由于T1NT比T2NT具有更好的分子平面性,其激发态下分子内电荷转移的程度较大,导致其在溶液中吸收光谱、发射光谱比T2NT呈现显著红移.冻结态下,分子平面性较差的T2NT显示出较短波长的发光.  相似文献   

6.
以1,3,5-三甲苯为原料,用硝酸和醋酸的混酸在5℃下乙酸酐溶液中对其硝化生成2,4,6-三甲基硝基苯(1),收率为90.6%。化合物1由铁粉和盐酸在水、乙醇混合溶剂中进行还原得到2,4,6-三甲基苯胺(2),收率为80.27%。将化合物2制备为盐酸重氮盐(3)后滴加到新制的甲醛肟中反应,浓缩得到2,4,6-三甲基苯甲醛(4),收率为52%。目标化合物结构经质谱、核磁氢谱确证,总收率37.82%,纯度98%。  相似文献   

7.
采用裸露簇和嵌入簇模型, 对β-MnO2 (001), (110), (111)三个晶面以及O2在(110)晶面的单址吸附模式(Pauling和Griffths模式), 进行从头计算. 从β-MnO2 (001), (110), (111)三个晶面的电子结构差异以及O2在(110)晶面吸附的吸附能、几何结构、集居数以及净电荷数分析得到: (001), (110), (111)三个晶面中(110)晶面的催化活性最高, 其活性顺序为(110)>(111)>(001). 氧气在(110)晶面的吸附, Pauling和Griffths两种吸附模式均存在, 属于化学吸附中的离子吸附. 氧气与MnO2固体间发生了单电子转移, 氧气得到电子被还原成O2-, 转移电子属于整个体系, 具有离域性.  相似文献   

8.
江洪  崔燕  方利  陈长水 《有机化学》2007,27(12):1590-1593
在甲苯溶液中,将N-硝基-2,4,6-三硝基苯胺与固体光气反应, 得到酰氯中间体, 该中间体无需分离, 然后直接与其它取代苯胺反应合成了13个未见文献报道的N-硝基-2,4,6-三硝基苯基脲类化合物, 其结构经1H NMR, IR, MS和元素分析确证. 初步测试结果表明: 当浓度为500 mg/L时, 部分化合物具有较好的抑制稗草生长活性.  相似文献   

9.
对氯苯基硫脲和2,4,6-三溴苯基硫脲的晶体结构和分子结构   总被引:1,自引:0,他引:1  
毛治华  单书香 《结构化学》1993,12(3):197-201
本文报道对氯苯基硫脲(Ⅰ)和2,4,6-三溴苯基硫脲(Ⅱ)的晶体结构和分子结构,(Ⅰ)属单斜晶系,空间群P2_1/a,a=10.671(1),b=8.912(1),c=9.128(1),β=106.85(1)°,Z=4,D_c=1.492g/cm~(-3),最后的偏离因子R=0.029,R_w=0.031;(Ⅱ)属正交晶系,空间群Pcab,a=9.997(2),b=7.876(4),c=27.920(3),Z=8,D_c=2.35g/cm~(-3),最后的偏离因子R=0.047,R_w=0.047。晶体结构测定结果表明,硫脲部分的四个原子组成平面三角形,苯环平面与硫脲平面的夹角分别为70.1°和80.3°。  相似文献   

10.
采用裸露簇和嵌入簇模型, 对β-MnO2 (001), (110), (111)三个晶面以及O2在(110)晶面的单址吸附模式(Pauling和Griffths模式), 进行从头计算. 从β-MnO2 (001), (110), (111)三个晶面的电子结构差异以及O2在(110)晶面吸附的吸附能、几何结构、集居数以及净电荷数分析得到: (001), (110), (111)三个晶面中(110)晶面的催化活性最高, 其活性顺序为(110)>(111)>(001). 氧气在(110)晶面的吸附, Pauling和Griffths两种吸附模式均存在, 属于化学吸附中的离子吸附. 氧气与MnO2固体间发生了单电子转移, 氧气得到电子被还原成O2-, 转移电子属于整个体系, 具有离域性.  相似文献   

11.
采用ReaxFF反应力场和分子动力学方法,研究了1,3,5-三氨基-2,4,6-三硝基苯(TATB)炸药晶体在沿不同方向冲击载荷下的滑移和各向异性。冲击方向分别垂直于(101)、(111)、(011)、(110)、(010)、(100)和(001)晶面,冲击强度为10 GPa。研究结果表明,各冲击方向下可能被激发的滑移系均在{001}面,而其它滑移系均因很大的剪切阻力不容易被激发,这与TATB晶体沿c轴的层状结构和平面分子结构相符。预测了七个冲击方向下最容易被激发的滑移系,分别为(101)/{001}100、(111)/{001}010、(011)/{001}010、(110)/{001}010、(010)/{001}110、(100)/{001}120和(001)/{001}010。TATB晶体的冲击响应具有各向异性,动力学过程中体系的应力、能量、温度和化学反应都依赖于冲击方向。对垂直于(100)和(001)晶面的冲击,体系在滑移过程中遭遇的剪切阻力较高、持续时间较长,使得能量和温度较快升高,化学反应较容易发生;对垂直于(101)和(111)晶面的冲击,体系在滑移过程中遭遇的阻力较小且出现次数少,使得能量和温度缓慢升高,化学反应不易发生;对其余冲击方向,体系的响应居中。据此评价了7个冲击方向的相对敏感程度:(101)、(111)(011)、(110)、(010)(100)、(001)。本研究有助于在微观层次深入认识动载荷下TATB的响应机制、结构与性能的关系,为高能低感炸药的设计和研制提供理论参考。  相似文献   

12.
By using the Amsterdam density functional (ADF) program based on first-principles, we have in detail investigated the Zn n (n = 2–20) clusters, several series of stacked structures, and the growth mechanism of zinc nanorods on silicone oil surfaces, which are isotropic, and can be generally considered as quasi-free sustained substrates. For the Zn n (n = 2–20) clusters, some new ground-state structures are found. The additional boundary atoms can enhance the stability of the stacked structures containing unit cells. The structures from the transverse array of the unit cells on (002) crystal plane are more stable than those stacked along the [002] direction as atomic number increases. The transverse arrayed structures can grow along the [002] direction. However, there is an atomic dislocation between the (002) crystal planes. In fact, the dislocation forms its one side surface of (101). By means of a single point SCF calculation in the ADF program, the adsorption energies per atom on the (101) and (002) crystal planes are calculated. According to the Boltzmann distribution probability law, the adsorption probability on the (101) plane is larger than that on the (002) plane, and the ratio between them is about 24–900. This shows that the Zn nanorods grow preferentially along the [110] direction on the (101) planes. These are in good agreement with our experimental findings.  相似文献   

13.
Platinum single crystal basal planes consisting of Pt(111), Pt(100), Pt(110) and reference polycrystalline platinum Pt(poly) were subjected to various potentiodynamic and potentiostatic electrochemical treatments in 0.1 M HClO4. Using the scanning flow cell coupled to an inductively coupled plasma mass spectrometer (SFC-ICP-MS) the transient dissolution was detected on-line. Clear trends in dissolution onset potentials and quantities emerged which can be related to the differences in the crystal plane surface structure energies and coordination. Pt(111) is observed to have a higher dissolution onset potential while the generalized trend in dissolution rates and quantities was found to be Pt(110)>P(100)≈Pt(poly)>Pt(111).  相似文献   

14.
The morphological evolution of uniform Cu(2)O nanocrystals with different morphologies in a weak acetic acid solution (pH = 3.5) has been studied for cubic, octahedral, rhombic dodecahedral, {100} truncated octahedral, and {110} truncated octahedral nanocrystals. Cu(2)O nanocrystals undergo oxidative dissolution in weak acid solution, but their morphological changes depend on the exposed crystal planes. We found that the stability of Cu(2)O crystal planes in weak acid solution follows the order of {100} ? {111} > {110} and determines how the morphology of Cu(2)O nanocrystals evolves. The stable {100} crystal planes remain, and new {100} facets form at the expense of the less stable {111} and {110} crystal planes on the surface of Cu(2)O nanocrystals. Density functional theory calculations reveal that the Cu-O bond on Cu(2)O(100) surface has the shortest bond length. These results clearly exemplify that the morphology of inorganic crystals will evolve with the change of local chemical environment, shedding light on fundamentally understanding the morphological evolution of natural minerals and providing novel insights into the geomimetic synthesis of inorganic materials in the laboratory.  相似文献   

15.
选择Fe1-xO基和Fe3O4基氨合成熔铁催化剂ZA-5和A110, 采用XRD原位反应器模拟真实的还原条件进行原位XRD实验. 通过对还原过程的物相跟踪分析和对前驱体、活性相的XRD微结构动态演化分析表明, ZA-5和A110的还原温度区间分别为300~362 ℃和343~450 ℃, 前者比后者具有更快的还原速度及更低的还原温度; ZA-5和A110在(211)和(110)方向的晶粒度比值(D(211)/D(110))分别为0.7014和0.8631, ZA-5催化剂具有更好的高活性(211)晶面, 其活性相微观应力明显比A110大, 且随着温度升高逐渐减小; 采用Rietveld全谱拟合结构分析和Popa模拟技术对活性相晶形进行模拟, 得到A110活性相晶形为凹陷立方体, ZA-5为立方体和球形的混合晶形, ZA-5的高活性晶面(111)和(211)晶面比A110生长更好; ZA-5催化剂活性相晶形从低温的类八面体向高温球形演化, 在355 ℃时有比较完善的八面体晶形, 此温度下(111)晶面生长得最好; ZA-5催化剂活性相与前驱体的八面体晶形结构匹配性好, 比A110更容易还原.  相似文献   

16.
A new strategy of using pre-grown crystals to study preferential adsorption of various additives is demonstrated for the electrocrystallization of Cu2O. In this method, micron-size Cu2O crystals with well-defined cubic and octahedral shapes were first electrochemically grown, and their crystallization was resumed in a medium containing the additive to be investigated (e.g., Na+, NH4+, SO42-, Cl-, dodecyl sulfate). This method makes it possible to systematically study the interaction of additives with specific planes (e.g., {100} of a cube and {111} of an octahedron) already present. By observing shape transformation over time, the relative stabilities of {100}, {111}, and {110} planes of Cu2O in various growth media could be determined. During this study, a general scheme of forming new crystal shapes containing crystallographic planes that cannot be directly stabilized by preferential adsorption alone was also established (i.e., rhombicuboctahedral shape of Cu2O containing {110} planes). This method can be extended to other crystal systems, which will enable us to classify common features of additives (e.g., charges, type of atoms) and crystallographic planes (e.g., atomic arrangement, surface termination, surface charge) required to allow for strong preferential adsorption.  相似文献   

17.
Molecular dynamics simulations of supported shock waves (shock pressure P(s) ~ 15 GPa) propagating along the [110], [011], [101], and [111] directions in crystalline nitromethane initially at T = 200 K were performed using the nonreactive Sorescu-Rice-Thompson force field [D. C. Sorescu, B. M. Rice, and D. L. Thompson, J. Phys. Chem. B 104, 8406 (2000)]. These simulations, combined with those from a preceding study of shocks propagating along [100], [010], and [001] directions in nitromethane for similar conditions of temperature and shock pressure [L. He, T. D. Sewell, and D. L. Thompson, J. Chem. Phys. 134, 124506 (2011)], have been used to study the post-shock relaxation phenomena. Shocks along [010] and [101] lead to a crystal-crystal structure transformation. Shocks propagating along [011], [110], [111], [100], and [001] exhibit plane-specific disordering, which was characterized by calculating as functions of time the 1D mean square displacement (MSD), 2D radial distribution function (RDF), and 2D orientation order parameter P(2)(θ) in orthogonal planes mutually perpendicular to the shock plane; and by calculating as functions of distance behind the shock front the Cartesian components of intermolecular, intramolecular, and total kinetic energies. The 2D RDF results show that the structural disordering for shocks along [100], [110], and [111] is strongly plane-specific; whereas for shocks along [001] and [011], the loss of crystal structural order is almost equivalent in the orthogonal planes perpendicular to the shock plane. Based on the entire set of simulations, there is a trend for the most extensive disordering to occur in the (010) and (110) planes, less extensive disordering to occur in the (100) plane, and essentially no disordering to occur in the (001) plane. The 2D P(2)(θ) and 1D MSD profiles show, respectively, that the orientational and translational disordering is plane-specific, which results in the plane-specific structural disordering observed in the 2D RDF. By contrast, the kinetic energy partitioning and redistribution do not exhibit plane specificity, as shown by the similarity of spatial profiles of the Cartesian components of the intermolecular, intramolecular, and total kinetic energies in orthogonal planes perpendicular to the shock plane.  相似文献   

18.
Hydroxide adsorption on the (111), (110), and (100) faces of silver electrodes from mixed NaOH/NaF solution is studied using cyclic voltammetry and in situ second harmonic generation (SHG). Cyclic voltammograms for the three low index silver planes in alkaline electrolytes are for the first time compared. They show two pairs of anodic and cathodic peaks in the potential interval below the equilibrium Ag/Ag(2)O potential. These are attributed to the specific adsorption of hydroxide ions followed by submonolayer oxide formation. The differences in the cyclic voltammograms for the (111), (110), and (100) planes are attributed to different (i) work functions, (ii) surface atomic densities, and (iii) corrugation potentials for these surfaces. Ex situ low energy electron diffraction (LEED) and reflection high energy electron diffraction (RHEED) show that disordered adlayers are formed on Ag(111) and Ag(100), in contrast to Ag(110), where ordered structures are produced in the region of the first pair of current peaks. In the region of the second pair of peaks, LEED indicates disordered oxide phases on each crystal plane and RHEED shows the presence of small islands of c(2 x 2) structure at some potentials on (110) and (100). SHG measurements were performed (i) in the potential scan mode at constant rotational angle and (ii) at constant potential as a function of the rotational angle. The isotropic (for the (111), (110), and (100) planes) and anisotropic (for the (110) and (111) planes) contributions to the SHG intensity were calculated by fitting the experimental data and are discussed in terms of their dependence on the charge density at the interface, on hydroxide adsorption, and on submonolayer oxide formation.  相似文献   

19.
A wide-angle x-ray diffraction (WAXD) study of the development of molecular orientation in the crystalline phase of ultra-high–molecular weight polyethylene films prepared by the gelation–crystallization method is presented. WAXD scans of the undrawn films show that the lamellae are oriented in the plane of the films. Upon drawing at 130°C, the orientation of the molecular chains changes from the direction normal to the film surface (ND) to the elongation direction. The decrease of the 200/020 intensity ratio at low draw ration (λ <10) indicates that double orientation develops during the transformation from the lamellar to the fibrillar morphology, with the a-axis oriented parallel to ND. The orientation distributions of the 110, 200, 020, and 002 planes of the orthorhombic unit cell of polyethylene were studied and characterized by the coefficients of a Legendre polynomial series. At a draw ratio of 4.5, the second-order coefficient, 〈P2(cos χ〉, already gets close to its limiting value, but it is shown that higher order coefficients of the polynomial series can be used to describe the evolution of the orentation, even up to λ = 50. The coefficients relative to the molecular chain orientation, 〈Pn(cos χ)〉c, can be calculated from different crystalline reflections. Curve-fitting calculations were made in order to improve the correlation between the results obtained from the orientation distribution of the 110, 020, and 002 planes. A Person VII function was found to give a better fit of the experimental curves than Gaussian or Lorentzian equations. © 1993 John Wiley & Sons, Inc.  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