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 546 毫秒
1.
刘蓓  廉源会  李智  陈光进 《化学学报》2014,72(8):942-948
选用了三种bio-MOFs (bio-MOF-1,bio-MOF-11,bio-MOF-100)材料,采用蒙特卡罗和分子动力学模拟研究了布洛芬分子在bio-MOF材料中的吸附和扩散性质. 结果发现,bio-MOF材料结构对药物分子布洛芬的吸附及扩散有很大影响. 其中,孔径控制着客体分子布洛芬的进入及扩散; 孔隙率大小与布洛芬的吸附量及自扩散系数大小成正比. 静电作用力对布洛芬分子的吸附有较小的促进作用. 另外还研究了布洛芬分子在MOF材料中的最佳吸附位及最优构型,发现布洛芬分子优先吸附在金属角落处,以及不同材料其吸附的布洛芬分子最优构型是不一样的.  相似文献   

2.
吴选军  郑佶  李江  蔡卫权 《物理化学学报》2013,29(10):2207-2214
采用优化的DREIDING力场参数, 通过巨正则系综蒙特卡洛(GCMC)模拟方法对H2在IRMOF-1、IRMOF-61和IRMOF-62共3种金属有机骨架(MOFs)材料中的吸附平衡性能进行了比较研究. 结果表明, 该力场能够在全压力范围内很好地复制H2在IRMOF-62材料中的等温吸附曲线; 但对低压下H2在IRMOF-61中的等温吸附曲线预测出现低估. 与IRMOF-1相比, 具有互穿骨架结构的IRMOF-61和IRMOF-62材料在常温下的储氢能力并无明显提高. 进一步比较77 K时100 kPa、3.0 MPa下H2在上述MOFs材料中达到吸附平衡时的几率密度分布发现, H2会优先吸附在Zn4O骨架附近靠近苯环的位置;对具有互穿结构的MOFs材料而言,由于其孔腔尺寸缩小, 使得H2优先吸附位区域零散化. 适当长度的有机配体形成的互穿骨架结构能增强与H2分子之间的相互作用, 具备较高的储氢能力; 而有机配体尺寸过长则会增加骨架结构中H2吸附死角, 对H2的吸附能力反而出现下降.  相似文献   

3.
随着计算机科学技术的飞速发展,理论计算特别是分子动力学模拟技术在研究受限流体的性质时发挥着独特的作用.本文综述了近年来水和醇类分子及其混合物在纳米孔道材料中传输扩散的研究进展,包括单组分水、甲醇和乙醇等在多种纳米孔道材料中的传输扩散,以及甲醇/水和乙醇/水等混合物在碳纳米管和沸石膜中的吸附和分离,讨论了体系温度、分子浓度以及纳米孔道材料结构等因素对水和醇类分子传输扩散过程的影响.  相似文献   

4.
场发射在扫描电子显微镜、平面显示器、压力传感器、加速度传感器以及电子束可寻址记忆器件等许多领域中得到了广泛的应用,分子基材料由于其结构和能带可设计,性质可调和柔性易加工等显著特点,被认为是新一代的场发射材料。本文综述了近年来分子基材料聚集态结构的场发射性质研究的新进展,特别是分子基材料的结构和聚集态形貌和尺寸对场发射性质的影响以及通过对分子基材料的杂化优化其场发射的性质,展望了分子基材料聚集态结构场发射的应用前景和发展趋势。  相似文献   

5.
利用分子模拟的方法研究了微量光气(COCl2)在微孔材料中的吸附和扩散性能, 并分析了材料结构的影响. 结果表明, 光气在金属有机框架材料(MOF)和共价有机框架材料(COF)中的吸附等温线主要表现为第Ⅰ类型和第Ⅴ类型吸附. 当光气压力较低时, COF材料和含有开放金属位点的材料对其吸附性能较好. 通过对不同压力下吸附量的比较发现, 吸附达到饱和前, 随着压力和孔隙率(VF)的升高, 材料对光气的吸附量增大. 通过分子动力学模拟研究光气在微孔材料中的扩散性能发现, 较强吸附位点的存在不利于光气在孔道中的扩散. 通过气体分子在材料中的径向函数分布图及模拟轨迹分析发现, 分子协同效应和空间位阻效应相互竞争决定了扩散速率的快慢. 综合评价材料的吸附和扩散性能发现, COF-102, COF-300, ZnMOF-74, Zn-DOBDC和PCN-60是理想的吸附材料, 这些材料可以应用于环境中光气泄漏的防治.  相似文献   

6.
利用分子模拟方法研究了十六烷基三甲基溴化铵(C16TAB)分子数对C16TAB/GO插层复合物的结构变化,探讨了C16TAB在GO层间的排列方式,并通过实验数据进行验证。模拟结果表明,优化后GO结构模型的层间距为0.849 nm;C16TAB/GO插层复合物的层间距随着C16TAB分子数的增加呈5个阶梯状逐渐增大,层间距分别为1.56、1.98、2.33、2.76和3.40 nm,插层饱和时C16TAB分子达到28个。实验结果显示,随着C16TAB分子数的增加,C16TAB/GO插层复合物的层间距逐渐增大,插层饱和时为3.40 nm,实验结果与模拟结果能够很好地吻合。C16TAB在GO层间可能的排列方式为1~5层平躺排列或单层平躺、单层倾斜和单层直立,从能量和结构的角度探明了C16TAB在GO层间的最优排列为1~5层平躺排列。  相似文献   

7.
MCM-22型分子筛中苯分子吸附行为的分子动力学模拟   总被引:3,自引:0,他引:3  
用分子动力学方法研究了纯硅MCM-22型分子筛中苯分子的吸附行为。计算结果表明,模拟采用刚性骨架或柔性骨架对苯分子的扩散系数并没有大的差别,这表明在较低吸附值的情况下,分子筛骨架的柔性对苯分子吸附和扩散并没有产生大的影响。  相似文献   

8.
水和盐分子在反渗透膜内扩散过程的分子模拟   总被引:6,自引:2,他引:4  
采用分子动力学模拟方法研究了水和盐分子(NaCl, MgCl2, CaSO4, K2SO4)在8种反渗透复合膜中的扩散状态及扩散系数. 并对膜材料的结构单体与水和盐分子在膜内扩散系数相关性进行了分析与讨论. 在所模拟的8种膜内, 随着膜种类的不同, 水分子在其中的扩散系数有明显的变化, 且扩散系数的变化规律与实验所得到的膜的水通量一一对应. NaCl分子中的Na+和Cl-在膜内的扩散速率不一致, 其扩散系数值在同种膜中相差较大, 且当盐分子单独存在时, 制约其在膜内扩散过程的离子只与膜种类有关, 与盐分子本身无关. 在同种膜中, 水分子的扩散过程不受体系中盐分子类型的影响.  相似文献   

9.
采用巨正则系综蒙特卡罗(GCMC)模拟方法, 对二氧化碳在5种具有相同拓扑结构的金属-有机骨架材料(IRMOFs), 即IRMOF-1, -8, -10, -14, -16中吸附产生的阶梯现象进行了详细的研究. 结果表明: 低温条件下, 孔径越大的IRMOFs越容易发生阶梯现象; 发生阶梯现象的转变压力与能够发生阶梯现象的转变温度都与孔径呈线性关系. 此外, 模拟结果进一步验证了二氧化碳分子之间的静电作用力是阶梯现象发生的关键因素. 这些规律将为金属-有机骨架材料(MOFs)的设计和改性以及二氧化碳在混合气体中的吸附分离提供有用的信息.  相似文献   

10.
金属-有机骨架材料中吸附气体的扩散速率   总被引:1,自引:0,他引:1  
采用分子动力学方法,以甲烷为探针分子研究了不同压力条件下气体在具有不同孔道结构的金属-有机骨架材料(MOFs)中的扩散速率.通过计算气体在八种材料中的自扩散系数,并结合气体分子在材料中的质心分布图等,讨论了气体扩散速率与孔道结构之间的关系.研究结果表明:对于同时含有孔笼(pocket)和三维正交孔道(channel)结构的MOF材料(P-C材料),低压时甲烷气体吸附在孔笼结构中,随着压力的升高,气体分子开始进入正交孔道,同时其自扩散系数增加;而对于只含有三维立方孔道结构的IRMOF(isoreticular MOF)系列材料,在中低压范围内,气体分子在其中的自扩散系数随压力变化较小.当压力进一步升高时,气体分子在材料孔道中的吸附逐渐接近饱和,其自扩散系数均降低.因此,在不同MOF材料中气体分子扩散速率的差异主要取决于孔道结构的不同.对P-C材料,中低压下通过控制压力可以控制气体在其中的扩散速率,从而为MOF材料在气体存储、分离等方面的实际应用提供参考信息.  相似文献   

11.
Interactions between alkali‐metal azides and metal–organic framework (MOF) derivatives, namely, the first and third members of the isoreticular MOF (IRMOF) family, IRMOF‐1 and IRMOF‐3, are studied within the density functional theory (DFT) paradigm. The investigations take into account different models of the selected IRMOFs. The mutual influence between the alkali‐metal azides and the π rings or Zn centers of the involved MOF derivatives are studied by considering the interactions both of the alkali‐metal cations with model aromatic centers and of the alkali‐metal azides with distinct sites of differently sized models of IRMOF‐1 and IRMOF‐3. Several exchange and correlation functionals are employed to calculate the corresponding interaction energies. Remarkably, it is found that, with increasing alkali‐metal atom size, the latter decrease for cations interacting with the π‐ring systems and increase for the azides interacting with the MOF fragments. The opposite behavior is explained by stabilization effects on the azide moieties and determined by the Zn atoms, which constitute the inorganic vertices of the IRMOF species. Larger cations can, in fact, coordinate more efficiently to both the aromatic center and the azide anion, and thus stabilizing bridging arrangements of the azide between one alkali‐metal and two Zn atoms in an η2 coordination mode are more favored.  相似文献   

12.
Increasing the resistance to humid environments is mandatory for the implementation of isoreticular metal–organic frameworks (IRMOFs) in industry. To date, the causes behind the sensitivity of [Zn44‐O)(μ‐bdc)3]8 (IRMOF‐1; bdc=1,4‐benzenedicarboxylate) to water remain still open. A multiscale scheme that combines Monte Carlo simulations, density functional theory and first‐principles Born–Oppenheimer molecular dynamics on IRMOF‐1 was employed to unravel the underlying atomistic mechanism responsible for lattice disruption. At very low water contents, H2O molecules are isolated in the lattice but provoke a dynamic opening of the terephthalic acid, and the lattice collapse occurs at about 6 % water weight at room temperature. The ability of Zn to form fivefold coordination spheres and the increasing basicity of water when forming clusters are responsible for the displacement of the organic linker. The present results pave the way for synthetic challenges with new target linkers that might provide more robust IRMOF structures.  相似文献   

13.
Hierarchical IRMOF‐3 nanosheets were firstly fabricated by a simple reflux strategy and were then characterized through Fourier transform infrared spectroscopy, X‐ray diffraction, scanning electron microscopy, transmission electron microscopy and X‐ray photoelectron spectroscopy. They revealed a high fluorescence quantum yield (13.2%) and showed excellent selectivity and sensitivity for 2,4,6‐trinitrophenol (TNP) over a concentration range of 1–29 μM in aqueous solution. This work demonstrates that the facile fabrication method for hierarchical IRMOF‐3 nanosheets with favorable selectivity and sensitivity for TNP could produce a new point of view on novel metal–organic framework nanomaterials for on‐line detection of organic pollutants in water.  相似文献   

14.
A copper iminopyridine complex has been immobilized on to a metal–organic framework (MOF) through postsynthetic modification of IRMOF‐3. The modified MOFs were fully demonstrated by using a variety of methods, and the structural integrity of the modified MOFs has been confirmed by powder X‐ray diffraction (XRD). Furthermore, it was shown that the modified IRMOF‐3 can act as an efficient solid catalyst for the synthesis of 2‐aminobenzothiazoles via the reaction of 2‐iodoanilines with isothiocyanates in a heterogeneous manner. Moreover, the catalyst could be facilely separated from the reaction mixture and reused for six consecutive cycles without significant degradation in catalytic activity. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

15.
Molecular screening of metal-organic frameworks for CO2 storage   总被引:1,自引:0,他引:1  
We report a molecular simulation study for CO2 storage in metal-organic frameworks (MOFs). As compared to the aluminum-free and cation-exchanged ZSM-5 zeolites and carbon nanotube bundle, IRMOF1 exhibits remarkably higher capacity. Incorporation of Na(+) cations into zeolite increases the capacity only at low pressures. By variation of the metal oxide, organic linker, functional group, and framework topology, a series of isoreticular MOFs (IRMOF1, Mg-IRMOF1, Be-IRMOF1, IRMOF1-(NH2)4, IRMOF10, IRMOF13, and IRMOF14) are systematically examined, as well as UMCM-1, a fluorous MOF (F-MOF1), and a covalent-organic framework (COF102). The affinity with CO2 is enhanced by addition of a functional group, and the constricted pore is formed by interpenetration of the framework; both lead to a larger isosteric heat and Henry's constant and subsequently a stronger adsorption at low pressures. The organic linker plays a critical role in tuning the free volume and accessible surface area and largely determines CO2 adsorption at high pressures. As a combination of high capacity and low framework density, IRMOF10, IRMOF14, and UMCM-1 are identified from this study to be the best for CO2 storage, even surpass the experimentally reported highest capacity in MOF-177. COF102 is a promising candidate with high capacity at considerably low pressures. Both gravimetric and volumetric capacities at 30 bar correlate well with the framework density, free volume, porosity, and accessible surface area. These structure-function correlations are useful for a priori prediction of CO2 capacity and for the rational screening of MOFs toward high-efficacy CO2 storage.  相似文献   

16.
The adsorption and desorption of n-hexane over Zeolite-5A has been investigated as a function of loading using simultaneous Fourier transform infrared (FTIR)-temperature-programmed desorption (TPD) measurements. The TPD profiles show a second peak developing at lower temperatures when loading exceeds 16 hexane molecules per Zeolite-5A unit cell or two molecules per alpha-cavity of the Zeolite-5A structure. The infrared spectra rule out two types of adsorption sites as the origin of the two peaks in the TPD. Changes in the conformation of the adsorbed hexane as a function of loading and temperature were followed by monitoring the position of the methylene stretching modes in the infrared spectra. With increasing loading, the adsorbed hexane adopts a stretched trans conformation. These changes occur at loading levels below 12 hexane molecules per Zeolite-5A unit cell. No change is observed above this loading, ruling out any conformational change at loadings where the second peak is seen in the TPD. The second peak in the TPD arises, therefore, from a combination of steric repulsion and loss of translational entropy.  相似文献   

17.
We report the synthesis and characterization of three different ordered mesoporous materials, labeled MCM‐48, SBA‐155, and SBA‐16 type materials, which were functionalized with gold nanoparticles using three different strategies. The functionalization strategies can be categorized as (i) in situ growth of gold nanoparticles, (ii) template loading, and (iii) diffusion loading of prefabricated gold nanoparticles. Two different particle sizes were employed in the latter two strategies, 5 nm and 10 nm. For all mesoporous structures, functionalization strategies, and particle sizes attempted, the materials retained their long‐range order upon incorporation of nanoparticles. From the adsorption isotherms, incorporation of gold nanoparticles altered the pore structure of the mesoporous support of some of the SBA‐15 and SBA‐16 type materials, with the effect on incorporation on the pore structure being particle size dependent in most cases. The majority of gold nanoparticles were found to reside on the external surface of the materials regardless of substrate and functionalization strategy; however, for the in situ synthesis and the template loading strategies, a significant fraction of the particles was determined to reside within the pore system of the material. In situ growth resulted in the highest content of gold nanoparticles in the solid phase. The relative effectiveness in retaining gold nanoparticles in the solid phase for each functionalization strategy was determined to be, in descending order, in situ synthesis, template loading, and diffusion loading.  相似文献   

18.
A simple and fast way to measure proton self‐diffusion coefficients of small penetrant molecules in semicrystalline polymers is introduced. The approach takes advantage of the strong static gradient of a mobile single‐sided NMR sensor and it is demonstrated on PE samples with varying degrees of crystallinity fully saturated in either toluene or n‐hexane. The self‐diffusion coefficients were measured using the gradient stimulated echo sequence appended with a CPMG. It is also shown for the first time, with demonstration on PE plates several millimeter thick with different aging histories, that one‐dimensional profiles of self‐diffusion coefficients as a function of depth can be easily obtained.  相似文献   

19.
Room temperature synthesis of metal-organic frameworks (MOFs) has been developed for four well-known MOFs: MOF-5, MOF-74, MOF-177, and MOF-199. A new isoreticular metal framework (IRMOF), IRMOF-0, having the same cubic topology as MOF-5, has been synthesized from acetylenedicarboxylic acid using this method to accommodate the thermal sensitivity of the linker. Despite acetylenedicarboxylate being the shortest straight linker that can be made into an IRMOF, IRMOF-0 forms as a doubly interpenetrating structure, owing to the rod-like nature of the linker.  相似文献   

20.
It was possible to determine the maximum loading of salicylic acid adsorbed onto γ-alumina and kaolinite clay after exposure to salicylic acid dissolved in hexane by examination using diffuse reflectance infrared Fourier transform infrared spectroscopy (DRIFTS). The maximum surface loading of salicylic acid (which resisted washing with fresh hexane) on γ-alumina was four times that observed using water as a solvent (approximately 3.0 compared with 0.7 molecules/nm2). Washing the sample with water removed the organic which was in excess to the maximum level observed for samples prepared with aqueous solution. The spectra of samples prepared with a loading up to the maximum observed with aqueous solution showed no significant differences to those of samples where the organic had been adsorbed from hexane (with the same surface loading). New peaks were observed for loadings greater than 1 molecules/nm2, but the salicylic acid was still present as carboxylate (with no clear evidence for the carbonyl group). Salicylic acid adsorbed more readily to the surface of kaolinite from solution in hexane than from aqueous solution (up to maximum average loading of 2 molecules/nm2). Washing the samples with water removed the organic to a loading in the region of 0.2 molecules/nm2, independent of the initial loading. Salicylic acid was adsorbed onto kaolinite as the carboxylate. The findings indicate that uptake is mediated by a surface water layer even in the absence of bulk water.  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