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 203 毫秒
1.
高岭石-水体系中水分子结构的分子动力学模拟   总被引:1,自引:0,他引:1  
以Hendricks模型为初始结构, 利用CLAYFF力场对高岭石-水体系进行无晶体学限制的分子动力学模拟. 结果表明, 层间水有三种类型: I型类似于Costanzo提出的“洞水”分子, 其HH矢量(水分子中从一个氢原子位置指向另一个氢原子位置的方向矢量)平行于(001)平面, 而C2轴稍微倾斜于(001)面法线; II型类似于“连接水”, 一个氢氧键指向临近的层间四面体氧形成氢键, 另一个氢氧键与(001)面近似平行; III型水分子在层间近似保持为竖直状, 一个氢与层间四面体氧形成氢键, 而另一个氢与对面层的羟基氧形成氢键. 高岭石羟基氢沿(001)晶面法线的浓度曲线显示一部分羟基指向变为近似平行于(001)面, 羟基氧因此能够暴露出来与层间水分子氢形成氢键. 此外, 模拟中还观察到部分II型水分子氧偏离于层间的平均位置而更靠近四面体层, 这和Costanzo的实验结果一致, 可能是X射线谱图中(002)弱衍射峰出现的原因.  相似文献   

2.
采用密度泛函理论B3LYP方法,在B3LYP/6-31G(d)理论水平上,构建高岭石的层间团簇模型Si6Al6O42H42(层间距为0.844 0和1.000 0nm),并对高岭石层间及其与n(n=1~3)个水分子相互作用的团簇的各种性质进行研究,如优化的几何构型、电子密度、氢键、能量、NBO电荷分布、振动频率等.结果表明,随着水分子个数n(n=1~3)的增加,体系的能量逐渐降低.水分子通过多种类型的氢键插层于高岭石层间,其中水分子间的氢键强度最强,其次是水分子与铝氧层之间形成的氢键,再次是水分子与硅氧层之间的氢键;层间距随着插层分子的增多而增大,但高岭石层间的活性位点依然存在,且位置较插层前没有明显变化.  相似文献   

3.
杨微  李晓蕾  王长生 《物理化学学报》2015,31(12):2285-2293
使用高精度从头算方法(含基组重叠误差校正)计算了水团簇(H2O)n (n = 8, 10, 16, 20, 22, 24)中的所有二体、三体和四体作用能,分析了水团簇中的多体效应.研究表明,二体作用对体系总作用能的贡献高达70%以上,三体作用对总作用能的贡献可高达25%,四体作用在总作用能中所占比例不超过3%,五体及以上多体作用能在总作用能中所占比例更小,不超过0.5%.本文研究还表明,两个水分子间距小于0.68 nm的二体作用、三个和四个水分子中最近的两个水分子间距小于0.31 nm的三体和四体作用对体系总作用能的贡献高达99.4%.因此,以生物体系为对象的分子模拟方法应该具备准确地模拟两个水分子间距小于0.68 nm的二体作用、三个和四个水分子中最近的两个分子间距小于0.31 nm的三体和四体作用的能力.  相似文献   

4.
用分子动力学方法模拟室温下不同浓度的聚甲基乙烯基醚/水体系的微观溶剂化结构.得到的径向分布函数和氢键给体和受体距离分布表明,聚合物与水形成的氢键比水之间形成的氢键短约0.005nm.准氢键C—H…O的数目是范德华作用对的7.2%.我们发现,在各浓度下,水分子并不能均匀地分布在聚合物结构单元上,即使在很稀的溶液(3.3%,质量分数)中,仍然有10%左右的醚氧没有和水分子形成氢键.这说明在溶液中,不但高分子链间有紧密的接触,而且高分子链内的链段间也有紧密的接触,导致链上的一些醚氧不能和水分子有效地接触而形成氢键.准氢键随浓度的变化和氢键的变化趋势类似,但形成准氢键的结构单元数目与形成氢键的结构单元数目比值在0.2附近.文献上用动态DSC测量低分子量聚甲基乙烯基醚(PVME)水溶液的相转变焓发现,在浓度为30%左右有一转折,与本模拟所得出的在浓度为27%左右氢键和准氢键比例的转折相关,这给相转变焓的转折点提供了分子尺度的微观解释.另外,浓度小于54%的溶液中存在“自由水”,在86%的浓溶液中每个结构单元大约与1.56个水分子缔合.  相似文献   

5.
合成了草酸根(ox)桥联[Cu(tmen)]^2+单元(tmen为N,N,N’,N’-四甲基乙二胺)的一维多聚铜配合物.X-ray单晶衍射表明,该配合物由Cu(Ⅱ)和四个ox氧原子、tmen的两个氮原子配位形成畸变的八面体构型.Cu…Cu间距为0.56095(34)nm和0.56594(40)nm.水分子和草酸根间氢键可以使该配合物形成三维超分子结构.  相似文献   

6.
报道了[Sc(NO3)3(OH2)3].(18-冠-6)的合成及其晶体结构.Sc(III)离子同三个双齿配体硝酸根与三个水分子氧配位,构成九配位的配合物.配位多面体是稍有扭曲的单帽四方反棱柱.配位水分子的六个氢原子分别与上下两层冠醚环上的氧原子生成氢键,形成多层夹心分子缔合物.  相似文献   

7.
罗素蓉  张锦楠  石梅  李奇 《化学学报》2004,62(11):1014-1018,M003
合成了一种新颖的苯三甲酸 -水分子为主体晶格的四丁基铵氢键包合物 ,(C4H9) 4 N+ [1,3 ,5 C6H3 (COOH) 2 (COO-) ]·2H2 O ,并进行了X射线晶体结构分析 .晶体为单斜晶系 ,P2 1 /c空间群 ,a =1 3 62 1( 2 )nm ,b =1 3 691( 2 )nm ,c =1 60 94( 2 )nm ,β =10 9 665 ( 3 )° ,V =2 82 61( 7)nm3 ,Z =4,R1 =0 0 5 0 8,wR =0 12 17.包合物中 ,苯三甲酸阴离子之间通过 3位与 5位羧基形成的O—H…O分子间氢键键连成沿b轴无限延伸的锯齿形长链 .位于长链较大空隙处的单一氢键键连的水分子二聚体 ,同时与两酸根的羧基生成给体氢键 ,形成一条苯三甲酸阴离子 -水分子复合宽带 .位于宽带外侧的 1位羧基与相邻宽带的水分子生成的氢键 ,将这些宽带连成平行于 ( 0 0 1)面的层状氢键主体晶格 .四丁基铵阳离子层填充在相邻主体晶格的层间空隙处 ,构成“三明治”式夹层结构的包合物  相似文献   

8.
油纸复合介质中水分子扩散行为的分子动力学模拟   总被引:3,自引:0,他引:3  
对不同温度下水分子在油纸复合介质中的扩散行为进行了分子动力学模拟研究. 通过分析水分子与纤维素形成的氢键发现, 油中的水分子在模拟过程中会逐渐扩散到纤维素内并与之形成氢键, 而纤维素内的水分子则与纤维素形成氢键后被束缚于纤维素中. 通过分析水分子的扩散系数发现, 由于油和纤维素的极性不同, 使得水分子在油和纤维素两种单介质中的扩散行为有较大差别, 而在复合介质中的扩散系数受水分子在油和纤维素中的比例影响较大, 两者表现出很强的相关性. 水分子和两介质的相互作用与两介质的极性也存在很大的关系, 且不同温度下水分子与两介质的相互作用能和水分子在油和纤维素中的比例也表现出了较强的相关性. 不同温度下水分子的不同分布弱化了温度对扩散系数的影响.  相似文献   

9.
张霞  张强  赵东霞 《化学学报》2012,70(3):60-66
利用分子动力学模拟方法对纯水溶液的氢键转化动力学性质进行了深入的微观探讨,溶液中非氢键构型为寿命较短(0.1~0.2 ps)的过渡态构型,我们发现氢键交换通过两种过渡构型完成,氢键角度扭曲激发后与氢键第一壳层水分子沿路径1交换,氢键径向拉伸激发后与氢键第二壳层水分子沿路径2交换,过渡态路径的选择具有温度依赖性.氢键转化需在旧氢键受体氢键过量和新氢键受体氢键不足,同时满足交换反应空间结构要求下才能完成.氢键交换反应对水分子平动和转动行为起着决定作用.  相似文献   

10.
在醋酸/水体系的工业分离中,溶液中的氢键对分离效率有很大影响.本文采用两种第一性原理方法,即从头算分子动力学模拟(AIMD)和量子化学计算(QCC),对由单个醋酸和不同水分子所组成聚合体的氢键相互作用进行了研究,采用极化统一模型和自洽反应场模型计算得到了聚合体在水溶液中的热力学数据.从QCC计算的气相和水溶液中的聚合自由能表明六元环在两种状态下都为最优结构,热力学数据反映出的各种结构的相对稳定性与AIMD模拟的环分布符合得相当一致.研究表明,由于存在醋酸和水分子间的氢键作用,稀醋酸/水溶液中的醋酸分离要比在浓醋酸溶液中困难得多.  相似文献   

11.
Electrofreezing of confined water   总被引:1,自引:0,他引:1  
We report results from molecular dynamics simulations of the freezing transition of TIP5P water molecules confined between two parallel plates under the influence of a homogeneous external electric field, with magnitude of 5 V/nm, along the lateral direction. For water confined to a thickness of a trilayer we find two different phases of ice at a temperature of T=280 K. The transformation between the two, proton-ordered, ice phases is found to be a strong first-order transition. The low-density ice phase is built from hexagonal rings parallel to the confining walls and corresponds to the structure of cubic ice. The high-density ice phase has an in-plane rhombic symmetry of the oxygen atoms and larger distortion of hydrogen bond angles. The short-range order of the two ice phases is the same as the local structure of the two bilayer phases of liquid water found recently in the absence of an electric field [J. Chem. Phys. 119, 1694 (2003)]. These high- and low-density phases of water differ in local ordering at the level of the second shell of nearest neighbors. The results reported in this paper, show a close similarity between the local structure of the liquid phase and the short-range order of the corresponding solid phase. This similarity might be enhanced in water due to the deep attractive well characterizing hydrogen bond interactions. We also investigate the low-density ice phase confined to a thickness of 4, 5, and 8 molecular layers under the influence of an electric field at T=300 K. In general, we find that the degree of ordering decreases as the distance between the two confining walls increases.  相似文献   

12.
Neutron diffraction elucidates the structures of two-dimensional (2D) water layers (278 K) or 2D ice layers confined in an organic slit-shaped nanospace. The two-dimensional ice phases reported here consist of individual eight-membered rings or folded-chain segments (263 K) and condensed twelve-membered irregular rings (20 K). This is quite different from bulk or other 2D ice structures; the latter usually form hexagonal honeycomb lattices. Both low-temperature structures typically feature water molecules which are surrounded by two or three other water molecules. Neutron diffraction and thermochemical studies indicate a liquid-solid-phase transition around 277 K for two-dimensional D2O layers. A further solid-solid-phase transition occurs between 263 and 20 K.  相似文献   

13.
The results of a computer simulation of flat fractures with widths of 0.63, 1.25, and 2.5 nm filled with vapor molecules in a silver iodide crystals at 260 K were presented. The two-dimensional gas of molecules adsorbed on the walls was found to be strongly clustered. Before the pore was filled, its walls had been covered with a monomolecular water film with a characteristic hexagonal structure. The perpendicular growth of the film was hindered by the hydrophobicity of its surface; the adsorbed molecules were bonded with the walls by interactions with the ions of the second crystal layer to form a specific orientational molecular order in the region of contact with the wall. On the wall with silver cations, the molecular energy was lower and the entropy higher than on the wall with iodide anions; on the wall with a lower energy, the adsorption started earlier and was more active. In an extremely narrow pore having room for only one molecular layer, the monomolecular film consists of spots held on opposite walls; in each spot, the orientational molecular order is the one characteristic of the wall with which the spot is in contact.  相似文献   

14.
Ice crystallized below 200 K has the diffraction pattern of a faulty cubic ice, and not of the most stable hexagonal ice polymorph. The origin and structure of this faulty cubic ice, presumed to form in the atmosphere, has long been a puzzle. Here we use large-scale molecular dynamics simulations with the mW water model to investigate the crystallization of water at 180 K and elucidate the development of cubic and hexagonal features in ice as it nucleates, grows and consolidates into crystallites with characteristic dimensions of a few nanometres. The simulations indicate that the ice crystallized at 180 K contains layers of cubic ice and hexagonal ice in a ratio of approximately 2 to 1. The stacks of hexagonal ice are very short, mostly one and two layers, and their frequency does not seem to follow a regular pattern. In spite of the high fraction of hexagonal layers, the diffraction pattern of the crystals is, as in the experiments, almost identical to that of cubic ice. Stacking of cubic and hexagonal layers is observed for ice nuclei with as little as 200 water molecules, but a preference for cubic ice is already well developed in ice nuclei one order of magnitude smaller: the critical ice nuclei at 180 K contain approximately ten water molecules in their core and are already rich in cubic ice. The energies of the cubic-rich and hexagonal-rich nuclei are indistinguishable, suggesting that the enrichment in cubic ice does not have a thermodynamic origin.  相似文献   

15.
Water nanoparticles play an important role in atmospheric processes, yet their equilibrium and nonequilibrium liquid-ice phase transitions and the structures they form on freezing are not yet fully elucidated. Here we use molecular dynamics simulations with the mW water model to investigate the nonequilibrium freezing and equilibrium melting of water nanoparticles with radii R between 1 and 4.7 nm and the structure of the ice formed by crystallization at temperatures between 150 and 200 K. The ice crystallized in the particles is a hybrid form of ice I with stacked layers of the cubic and hexagonal ice polymorphs in a ratio approximately 2:1. The ratio of cubic ice to hexagonal ice is insensitive to the radius of the water particle and is comparable to that found in simulations of bulk water around the same temperature. Heating frozen particles that contain multiple crystallites leads to Ostwald ripening and annealing of the ice structures, accompanied by an increase in the amount of ice at the expense of the liquid water, before the particles finally melt from the hybrid ice I to liquid, without a transition to hexagonal ice. The melting temperatures T(m) of the nanoparticles are not affected by the ratio of cubic to hexagonal layers in the crystal. T(m) of the ice particles decreases from 255 to 170 K with the particle size and is well described by the Gibbs-Thomson equation, T(m)(R) = T(m)(bulk) - K(GT)/(R - d), with constant K(GT) = 82 ± 5 K·nm and a premelted liquid of width d = 0.26 ± 0.05 nm, about one monolayer. The freezing temperatures also decrease with the particles' radii. These results are important for understanding the composition, freezing, and melting properties of ice and liquid water particles under atmospheric conditions.  相似文献   

16.
A binary liquid mixture, containing the Lennard-Jones molecules A and B, in equilibrium with a bulk liquid reservoir near the point of phase separation, confined between atomistic chemically patterned walls, is studied by grand canonical Monte Carlo simulations. In the bulk, the B-rich phase is stable and the A-rich phase is metastable. The walls bear patches attractive to A; when the walls are close, A-rich liquid bridges condense between the patches. The normal and lateral forces on the walls are measured as a function of the wall separation and of the lateral displacement between the patches on opposite walls. When there are one or two molecular layers in the bridge and the wall lattice constant is close to that of crystalline A, the normal and lateral forces depend strongly on the registry of the wall lattices, varying in an oscillatory manner.  相似文献   

17.
Molecular dynamics (MD) simulations of water confined in nanospaces between layers of talc (system composition Mg(3)Si(4)O(10)(OH)(2) + 2H(2)O) at 300 K and pressures of approximately 0.45 GPa show the presence of a novel 2-D ice structure, and the simulation results at lower pressures provide insight into the mechanisms of its decompression melting. Talc is hydrophobic at ambient pressure and temperature, but weak hydrogen bonding between the talc surface and the water molecules plays an important role in stabilizing the hydrated structure at high pressure. The simulation results suggest that experimentally accessible elevated pressures may cause formation of a wide range of previously unknown water structures in nanoconfinement. In the talc 2-D ice, each water molecule is coordinated by six O(b) atoms of one basal siloxane sheet and three water molecules. The water molecules are arranged in a buckled hexagonal array in the a-b crystallographic plane with two sublayers along [001]. Each H(2)O molecule has four H-bonds, accepting one from the talc OH group and one from another water molecule and donating one to an O(b) and one to another water molecule. In plan view, the molecules are arranged in six-member rings reflecting the substrate talc structure. Decompression melting occurs by migration of water molecules to interstitial sites in the centers of six-member rings and eventual formation of separate empty and water-filled regions.  相似文献   

18.
The phase behavior of confined water is a topic of intense and current interest due to its relevance in biology, geology, and materials science. Nevertheless, little is known about the phases that water forms even when confined in the simplest geometries, such as water confined between parallel surfaces. Here we use molecular dynamics simulations to compute the phase diagram of two layers of water confined between parallel non hydrogen bonding walls. This study shows that the water bilayer forms a dodecagonal quasicrystal, as well as two previously unreported bilayer crystals, one tiled exclusively by pentagonal rings. Quasicrystals, structures with long-range order but without periodicity, have never before been reported for water. The dodecagonal quasicrystal is obtained from the bilayer liquid through a reversible first-order phase transition and has diffusivity intermediate between that of the bilayer liquid and ice phases. The water quasicrystal and the ice polymorphs based on pentagons are stabilized by compression of the bilayer and are not templated by the confining surfaces, which are smooth. This demonstrates that these novel phases are intrinsically favored in bilayer water and suggests that these structures could be relevant not only for confined water but also for the wetting and properties of water at interfaces.  相似文献   

19.
We report the results of a theoretical study of locally ordered fluctuations in a quasi-two-dimensional colloid fluid. The fluctuations in the equilibrium state are monitored by the aperture cross-correlation function of radiation scattered by the fluid, as calculated from molecular dynamics simulations of near hard spheres with diameter sigma confined between smooth hard walls. These locally ordered fluctuations are transient; their decay can be monitored as a function of the time between the cross-correlated scattered radiation signals, but only the single-time cross-correlated signals are discussed in this paper. Systems with thicknesses less than two hard sphere diameters were studied. For wall separation H in the range 1 sigma/=1.57 sigma, hexagonal fluctuations persist in the dense liquid up to H=1.75 sigma, and fluctuations with square ordered symmetry, that of the solid to which the liquid freezes, only emerge at densities approximately 2% below freezing. For H=1.8 sigma and 1.85 sigma, hexagonal ordered flucuations are no longer found, and the square ordered fluctuations dominate the dense liquid region as the system freezes into a two layer square solid. For H=1.9 sigma and 1.95 sigma, where the liquid freezes into a two layer hexagonal solid, both square and hexagonal ordered fluctuations are observed. At lower densities, the ordered fluctuations only exhibit square symmetry. Hexagonal ordered fluctuations appear at densities approximately 7% below freezing and become more dominant as the density is increased, but the square ordered fluctuations persist until the system is converted into the solid.  相似文献   

20.
This paper reports on a new method for the preparation of mesoporous silica membranes on alumina hollow fibers. A surfactant-silica sol is filled in the lumen of an alpha-alumina hollow fiber. The filtration technique combined with an evaporation-induced self-assembly (EISA) process results in the formation of a continuous ordered mesoporous silica layer on the outer side of alpha-alumina hollow fibers. X-ray diffraction (XRD), transmission electron microscopy (TEM), and nitrogen isothermal adsorption measurements reveal that these membranes possess hexagonal (P6mm) mesostructures with pore diameters of 4.48 nm and BET surfaces of 492.3 m(2) g(-1). Scanning electron microscopy (SEM) studies show that the layers are defect free and energy-dispersive spectroscopy (EDS) mapping images further confirm the formation of continuous mesoporous silica layer on the outer side of alpha-alumina hollow fibers. Nitrogen and hydrogen permeance tests show that the membranes are defect free.  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