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 265 毫秒
1.
甘油水溶液氢键特性的分子动力学模拟   总被引:3,自引:0,他引:3  
为了研究低温保护剂溶液的结构和物理化学特性, 以甘油为保护剂, 采用分子动力学方法, 对不同浓度的甘油和水的二元体系进行了模拟. 得到了不同浓度的甘油水溶液在2 ns内的分子动力学运动轨迹, 通过对后1 ns内运动轨迹的分析, 得到了各个原子对的径向分布函数和甘油分子的构型分布. 根据氢键的图形定义, 分析了氢键的结构和动力学特性. 计算了不同浓度下体系中平均每个原子(O和H)和分子(甘油和水)参与氢键个数的百分比分布及其平均值. 同时还计算了所有氢键、水分子之间的氢键以及甘油与水分子之间的氢键的生存周期.  相似文献   

2.
采用分子模拟方法研究了正交晶系冰晶(020)生长面在不同浓度甘油水溶液中的生长情况. 通过统计分析氢键数、 密度分布函数、 均方根偏差和原子间径向分布函数研究了水分子和甘油分子的动态行为. 结果表明, 甘油分子在水溶液中可与水分子形成大量氢键, 这使水分子间的氢键作用受到抑制, 降低了水分子的扩散性, 致使冰晶不易成核和生长; 另外, 一些甘油分子可代替水分子吸附在晶面上, 甚至占据晶格位点, 这种行为打破了冰晶的对称性并且降低了冰晶的生长速率. 因此, 甘油可同时在晶面和液相中抑制冰晶的生长.  相似文献   

3.
构建水滑石(LDHs-Cl-yH2O)周期性计算模型, 选用密度泛函理论-赝势平面波法对模型进行几何全优化, 从结构参数、Mulliken电荷布居、态密度(DOS)、能量等角度研究层间Cl-和水分子的分布形态以及与LDHs层板间的超分子作用. 计算结果表明, LDHs-Cl主客体间存在着较强的超分子作用, 主要包括静电和氢键作用. LDHs-Cl层间引入水分子后, 随着水分子数的增加, 层间距逐渐增大后趋于平衡. 水合过程中氢键作用比静电作用更占优势, layer-water型氢键要略强于anoin-water型氢键. 当y=1, 2时, Cl-与水分子所在平面以平行层板的方式存在于LDHs层板间, 并且与两层板的距离基本相等; 当y=3, 4时, Cl-与水分子则以偏向某一层的方式随机地存在于LDHs 层板间. 随着层间水分子增加, LDHs-Cl-yH2O由离子型晶体向分子型晶体转化, LDHs-Cl的水合具有饱和量.  相似文献   

4.
为研究不同结构的表面活性剂分子在溶液中胶束化能力的差异, 采用分子动力学方法模拟三种烷基芳基磺酸盐在真空和水溶液环境下的结构与相互作用. 利用自由能微扰(FEP)方法计算了水合自由能, 发现与用传统热力学表面张力法测定自制的烷基芳基磺酸盐结果一致. 研究表明: 烷基芳基磺酸盐在水溶液中的胶束化过程是自发进行的, 随着分子结构中芳环向长烷基链中间位置移动, 胶束化能力和胶束稳定性均下降; 疏水基周围水分子的“冰山结构”会影响胶束的稳定性, 而水分子中氢键的生存周期是反映冰山结构变化的重要指标; 同时, 亲水基与水分子间形成氢键的数目会增强或减弱分子脱离胶束体的趋势, 从而影响胶束结构的稳定性.  相似文献   

5.
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   

6.
采用拉曼光谱研究了水-乙腈-二甲基亚砜三元水溶液体系中氢键作用对分子结构的影响.结果表明,在三元水溶液中乙腈C≡N键的电子云向碳原子发生偏移,二甲基亚砜中S=O双键的电子云向硫原子发生偏移;在三元体系中乙腈和二甲基亚砜与水形成氢键时存在明显的竞争关系,出现乙腈分子和二甲基亚砜分子共用1个水分子形成复合物的情况,并且随着水含量的增加,共用水分子的情况逐渐消失.  相似文献   

7.
特殊缔合体系TFE水溶液分子动力学模拟   总被引:2,自引:0,他引:2  
三氟乙醇(TFE)水溶液是一类特殊的缔合体系. 采用分子动力学模拟方法结合核磁共振化学位移研究了TFE水溶液体系全浓度范围的氢键网络, 并对动力学模拟结果和核磁共振化学位移进行了比较. 从径向分布函数(RDF)发现, TFE水溶液中存在着强氢键, 而体系中的C—H…O弱相互作用较为明显, 也不能忽略. 氢键网络分析发现TFE 水溶液体系的氢键大致分为以下三个区域: 在水富集区域, 水分子倾向于自身缔合形成稳定的簇结构, 随着TFE 浓度的增加, 水的有序结构受到破坏, 水分子和TFE分子发生交叉缔合作用形成氢键; 在TFE富集区域, 水分子较少, TFE分子自身通过氢键形成多缔体结构. 此外, 分子动力学统计的平均氢键数的变化和文献报导的核磁共振化学位移变化趋势相同, 实验和理论的结果吻合较好.  相似文献   

8.
使用分子动力学研究了乙醇与水分子在纳米金管内按照不同比例混合时的吸附现象,并利用径向密度分布函数及水和乙醇分子所形成的平均氢键数来探讨纳米限制效应.结果表明,径向密度分布函数和氢键数目受纳米金管影响较大.另外,水与金管之间的作用力比乙醇与金管之间的大,导致水分子形成的平均氢键数不同于乙醇分子的.  相似文献   

9.
不同电解质体系水的拉曼谱的研究   总被引:4,自引:0,他引:4  
通过一系列电解质体系水的拉曼光谱测量,得到了阴、阳离子种类和浓度引起的水伸缩振动和弯曲振动谱带丰富的变化信息,ClO4^-能有效地破坏水分子间的氢键,随着ClO4浓度的增加,水分子间的氢键并非逐步被打断,而是氢键被破坏的水分子越来越多,从而使水分子有序度增大,这种氢键破坏方式符合水的混合模型(MixtureModel)SO^2-4浓度的增对水的Raman光谱影响较小,是由于SO^2-4与水分了间的  相似文献   

10.
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   

11.
In this study, we use molecular dynamics simulations to investigate and compare the interactions of DPPC bilayers with and without saccharides (glucose or trehalose) under dehydrated conditions. Results from the simulations indicate that unilamellar bilayers lose their structural integrity under dehydrated conditions in the absence of saccharides; however, in the presence of either glucose or trehalose, the bilayers maintain their stability. Hydrogen bond analysis shows that the saccharide molecules displace a significant amount of water surrounding the lipid headgroups. At the same time, the additional hydrogen bonds formed between water and saccharide molecules help to maintain a hydration layer on the lipid bilayer interface. On the basis of the hydrogen bond distributions, trehalose forms more hydrogen bonds with the lipids than glucose, and it is less likely to interact with neighboring saccharide molecules. These results suggest that the interaction between the saccharide and lipid molecules through hydrogen bonds is an essential component of the mechanism for the stabilization of lipid bilayers.  相似文献   

12.
The title compound, C18H19Cl2NO4·C6H12N2·H2O, is a cocrystal hydrate containing the active pharmaceutical ingredient felodipine and diazabicyclo[2.2.2]octane (DABCO). The DABCO and water molecules are linked through O—H...N hydrogen bonds into chains around 21 screw axes, while the felodipine molecules form N—H...O hydrogen bonds to the water molecules. The felodipine molecules adopt centrosymmetric back‐to‐back arrangements that are similar to those present in all of its four reported polymorphs. The dichlorophenyl rings also form π‐stacking interactions. The inclusion of water molecules in the cocrystal, rather than formation of N—H...N hydrogen bonds between felodipine and DABCO, may be associated with steric hindrance that would arise between DABCO and the methyl groups of felodipine if they were directly involved in hydrogen bonding.  相似文献   

13.
In the crystal structure of O,O′‐diethyl N‐(2,4,6‐trimethylphenyl)thiophosphate, C13H22NO2PS, two symmetrically independent thiophosphoramide molecules are linked through N—H…S and N—H…π hydrogen bonds to form a noncentrosymmetric dimer, with Z′ = 2. The strengths of the hydrogen bonds were evaluated using density functional theory (DFT) at the M06‐2X level within the 6‐311++G(d,p) basis set, and by considering the quantum theory of atoms in molecules (QTAIM). It was found that the N—H…S hydrogen bond is slightly stronger than the N—H…π hydrogen bond. This is reflected in differences between the calculated N—H stretching frequencies of the isolated molecules and the frequencies of the same N—H units involved in the different hydrogen bonds of the hydrogen‐bonded dimer. For these hydrogen bonds, the corresponding charge transfers, i.e. lp (or π)→σ*, were studied, according to the second‐order perturbation theory in natural bond orbital (NBO) methodology. Hirshfeld surface analysis was applied for a detailed investigation of all the contacts participating in the crystal packing.  相似文献   

14.
A molecular dynamics simulation of water molecules through a Au nanotube with a diameter of 20 A at bulk densities 0.8, 1, and 1.2 gcm(3) has been carried out. The water molecules inside a nanoscale tube, unlike those inside a bulk tube, have a confined effect. The interaction energy of the Au nanotube wall has a direct influence on the distribution of water molecules inside the Au tube in that the adsorption of the water molecules creates shell-like formations of water. Moreover, the high number of adsorbed molecules has already achieved saturation at the wall of the Au nanotube at three bulk densities. This work compares the distribution percentage profiles of hydrogen bonds for different regions inside the tube. The structural characteristics of water molecules inside the tube have also been studied. The results reveal that the numbers of hydrogen bonds per water molecule influence the orientational order parameter q. In addition, the phenomenon of a group of molecules bonded inside the tube can be observed as the number of hydrogen bonds increase.  相似文献   

15.
NMR J-couplings across hydrogen bonds reflect the static and dynamic character of hydrogen bonding. They are affected by thermal and solvent effects and can therefore be used to probe such effects. We have applied density functional theory (DFT) to compute the NMR (n)J(N,H) scalar couplings of a prototypical Chagas disease drug (metronidazole). The calculations were done for the molecule in vacuo, in microsolvated cluster models with one or few water molecules, in snapshots obtained from molecular dynamics simulations with explicit water solvent, and in a polarizable dielectric continuum. Hyperconjugative and electrostatic effects on spin-spin coupling constants were assessed through DFT calculations using natural bond orbital (NBO) analysis and atoms in molecules (AIM) theory. In the calculations with explicit solvent molecules, special attention was given to the nature of the hydrogen bonds formed with the solvent molecules. The results highlight the importance of properly incorporating thermal and solvent effects into NMR calculations in the condensed phase.  相似文献   

16.
Two polymorphs of bis(2‐carbamoylguanidinium) fluorophosphonate dihydrate, 2C2H7N4O+·FO3P2−·2H2O, are presented. Polymorph (I), crystallizing in the space group Pnma, is slightly less densely packed than polymorph (II), which crystallizes in Pbca. In (I), the fluorophosphonate anion is situated on a crystallographic mirror plane and the O atom of the water molecule is disordered over two positions, in contrast with its H atoms. The hydrogen‐bond patterns in both polymorphs share similar features. There are O—H...O and N—H...O hydrogen bonds in both structures. The water molecules donate their H atoms to the O atoms of the fluorophosphonates exclusively. The water molecules and the fluorophosphonates participate in the formation of R44(10) graph‐set motifs. These motifs extend along the a axis in each structure. The water molecules are also acceptors of either one [in (I) and (II)] or two [in (II)] N—H...O hydrogen bonds. The water molecules are significant building elements in the formation of a three‐dimensional hydrogen‐bond network in both structures. Despite these similarities, there are substantial differences between the hydrogen‐bond networks of (I) and (II). The N—H...O and O—H...O hydrogen bonds in (I) are stronger and weaker, respectively, than those in (II). Moreover, in (I), the shortest N—H...O hydrogen bonds are shorter than the shortest O—H...O hydrogen bonds, which is an unusual feature. The properties of the hydrogen‐bond network in (II) can be related to an unusually long P—O bond length for an unhydrogenated fluorophosphonate anion that is present in this structure. In both structures, the N—H...F interactions are far weaker than the N—H...O hydrogen bonds. It follows from the structure analysis that (II) seems to be thermodynamically more stable than (I).  相似文献   

17.

The crystals of the pyrazine-2,6-dicarboxylic acid dihydrate [C 4 H 2 N 2 (COOH) 2 ]·2H 2 O or H 2 (2,6-PZDC)] crystallize in the monoclinic system, space group C2/m. Their structure is composed of planar layers in which the acid and the water molecules interact via a network of hydrogen bonds. The layers are also hydrogen bonded. Hexaaquamagnesium(II) pyrazine-2,6-dicarboxylate [Mg(H 2 O) 6 ] 2+ [C 4 H 2 N 2 (COO) 2 ] 2 m crystallizes in the monoclinic system, space group P2 1 / n . The magnesium(II) cation is surrounded by six water molecules located at the apices of an almost regular octahedron with the mean Mg-O bond distance of 2.068 Å. The 2,6-PZDC anions are planar and are acceptors in a network of hydrogen bonds donated by the coordinated water molecules.  相似文献   

18.
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   

19.
随机产生单笼形水分子簇(H2O)n(n=8~36),经分类统计后发现,在笼形水分子簇中,其1221,1212,2121和2112四类氢键的个数与水分子和氢键总数之间有定量关系,且1212类氢键的个数与2121类的氢键始终相等.如果笼形水分子簇中某一类氢键数已知,则它的其余三类氢键的个数也随即确定.  相似文献   

20.
Molecular dynamics simulations are used to investigate the interaction of the sugars trehalose, maltose, and glucose with a phospholipid bilayer at atomic resolution. Simulations of the bilayer in the absence or in the presence of sugar (2 molal concentration for the disaccharides, 4 molal for the monosaccharide) are carried out at 325 and 475 K. At 325 K, the three sugars are found to interact directly with the lipid headgroups through hydrogen bonds, replacing water at about one-fifth to one-quarter of the hydrogen-bonding sites provided by the membrane. Because of its small size and of the reduced topological constraints imposed on the hydroxyl group locations and orientations, glucose interacts more tightly (at identical effective hydroxyl group concentration) with the lipid headgroups when compared to the disaccharides. At high temperature, the three sugars are able to prevent the thermal disruption of the bilayer. This protective effect is correlated with a significant increase in the number of sugar-headgroups hydrogen bonds. For the disaccharides, this change is predominantly due to an increase in the number of sugar molecules bridging three or more lipid molecules. For glucose, it is primarily due to an increase in the number of sugar molecules bound to one or bridging two lipid molecules.  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