首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到18条相似文献,搜索用时 281 毫秒
1.
铅基复合钙钛矿型弛豫铁电单晶体PMNT的生长基元为多种[BO~6]配位八面体,晶体生长过程可视为多种八面体基元与Pb^2^+的组装过程。这些生长基元向{111}面叠合时易采取法向生长机制,向{001}面叠合时易采取层状生长机制,由此决定了晶体生长速度的各向异性与晶体的形貌。Bridgman法生长的PMNT晶体在生长过程中由内向生长机制形成规则的负晶结构;在晶体生长过程中,在其自然表面上可形成正形与负形两种形貌;在高温退火过程中,由于PbO的分解,晶体表面上可形成类似"蚀象"的构型,这些可从[BO~6]八面体生长基元的组装或拆分方面获得解释。  相似文献   

2.
铅基复合钙钛矿型弛豫铁电单晶体PMNT的生长基元为多种[BO_6]配位八面体,晶体生长过程可视为多种八面体基元与Pb~(2 )的组装过程.这些生长基元向{111}面叠合时易采取法向生长机制,向{001}面叠合时易采取层状生长机制,由此决定了晶体生长速度的各向异性与晶体的形貌.Bridgman法生长的PMNT晶体在生长过程中由内向生长机制形成规则的负晶结构;在晶体生长过程中,在其自然表面上可形成正形与负形两种形貌;在高温退火过程中,由于PbO的分解,晶体表面上可形成类似“蚀象”的构型,这些可从[BO_6]八面体生长基元的组装或拆分方面获得解释.  相似文献   

3.
铅基复合钙钛矿型弛豫铁电单晶体PMNT,PZNT的生长基元为多种[BO~6]配位八面体。这些同型生长基元受本身稳定性的制约而在熔体中存在的几率不同。相对于[MgO~6]^1^0^-,[ZnO~6]^1^0^-八面体基元来说,[NbO~6]^7^-,[TiO~6]^8^-是更为有利的八面体基元。在基元组装过程中,各种[BO~6]八面体基元在稳定性、尺寸大小与电价上的分异致使生长界面对基元有一定的选择性,从而造成了晶体生长时成分与结构的短程起伏,并为有序畴及其它化学缺陷团簇的形成提供了条件。当加入掺质PbTiO~3时,由于[TiO~6]^8^-与[NbO~6]^7^-两种基元在组装时的类聚性及[TiO~6]^8^-对晶体稳定性的贡献,晶体的微区成分与结构得以调制,焦发石相得以抑制,这构成了用Bridgman法能直接从熔体中生长出纯钙钛矿相PMNT单晶的基础。而[MgO~6]^1^0^-与[ZnO~6]^1^0^-八面体基元的差致使PMNT,PZNT两单晶的生长难度有别,这在选择合适的生长方法时需加以考虑。  相似文献   

4.
铅基复合钙钛矿型弛豫铁电单晶体PMNT,PZNT的生长基元为多种[BO~6]配位八面体。这些同型生长基元受本身稳定性的制约而在熔体中存在的几率不同。相对于[MgO~6]^1^0^-,[ZnO~6]^1^0^-八面体基元来说,[NbO~6]^7^-,[TiO~6]^8^-是更为有利的八面体基元。在基元组装过程中,各种[BO~6]八面体基元在稳定性、尺寸大小与电价上的分异致使生长界面对基元有一定的选择性,从而造成了晶体生长时成分与结构的短程起伏,并为有序畴及其它化学缺陷团簇的形成提供了条件。当加入掺质PbTiO~3时,由于[TiO~6]^8^-与[NbO~6]^7^-两种基元在组装时的类聚性及[TiO~6]^8^-对晶体稳定性的贡献,晶体的微区成分与结构得以调制,焦发石相得以抑制,这构成了用Bridgman法能直接从熔体中生长出纯钙钛矿相PMNT单晶的基础。而[MgO~6]^1^0^-与[ZnO~6]^1^0^-八面体基元的差致使PMNT,PZNT两单晶的生长难度有别,这在选择合适的生长方法时需加以考虑。  相似文献   

5.
在热液条件下晶体的生长基元与晶体形成机理   总被引:36,自引:0,他引:36  
本文通过电泳实验研究了热液条件下,水晶(SiO_2)、钛酸钡(BaTiO_3)晶体生长基元与晶体形成机理。提出晶体生长基元具有负离子配位多面体结构,与晶体中负离子配位结构相当。根据负离子多面体往晶体各族晶面上叠合的难易程度,解释了晶体结晶形貌与物理、化学条件之间的关系。随着生长条件的变化生长基元的维度也有所不同,不同维度的生长基元往各族晶面上的叠合速率发生相应改变,从而解释了晶体形貌的多变性。  相似文献   

6.
β-BBO晶体的结晶习性与形成机理   总被引:5,自引:0,他引:5  
从结晶化学角度出发,研究了β-BBO结构中的基本结构单元,[B_3O_6]~(3-)环状络阴离子的结晶方位与晶体各族晶面的对应关系,根据Na_2O-BaB_2O_4溶液结构的测定资料,提出β-BBO晶体生长基元的结构形式和生长基元往各族晶面上叠合的规律。讨论了β-BBO晶体结晶习性的形成机制。由于物理化学条件的不同,生长基元的维度也不相同,而不同维度的生长基元往晶体各族晶面上的叠合速率比也会发生相应的变化,这是导致晶体形貌上形成多变性的原因。  相似文献   

7.
晶体生长基元与晶体结晶习性   总被引:7,自引:0,他引:7  
仲维卓  唐鼎元 《结构化学》1995,14(5):463-468
本文从结晶化学角度出发,结合晶体生长实际研究了AO2、A2O3和ABO3型晶体中AO6八面体的结晶方位和晶体方位和晶体结晶形貌之间的关系,通过BBO、LBO高温溶液结构的测定和对水热条件下SiO2和BTO溶液结构的分析,运用负离子配位多面体生长基元的理论模型解释了晶体的生长形态。  相似文献   

8.
几种极性有机晶体的生长习性与形成机理   总被引:4,自引:0,他引:4  
有机晶体特别是极性有机晶体, 在不同的溶剂中具有明显不同的生长习性。本文通过对4-氨基-4'-硝基二苯硫醚(ANDS)等几种典型极性有机晶体在不同溶剂中的生长习性和结晶形貌的讨论, 提出了偶极生长基元叠合模型, 从两个方面探讨了这些习性的形成机理, 即(1)有机晶体在不同的溶剂中具有不同结构和形式的生长基元(对于极性有机晶体而言, 这些生长基元都具有偶极子特征), 而不同的生长基元往晶体的各个面族上叠合的相对速率不同, 从而导致了晶体习性的改变; (2)晶体生长界面的性质不同, 特别是对于极性晶体, 晶体界面的极性不同;不同的溶剂与生长晶体的界面相互作用不同, 即使同种溶剂对晶体不同界面上的作用也不同, 因而改变了生长界面的性质,影响了生长基元在晶体界面, 特别是晶体正、负极面上的叠合速率, 从而导致了晶体形貌的变化。由此比较合理地解释了晶体所呈现的不同的生长习性, 特别是合理地解释了极性有机晶体所呈现的极性生长特征。  相似文献   

9.
龙英才  田正荣 《化学学报》1988,46(4):309-314
本文研究了DEA(二乙醇胺)-Na2O-SiO2-Al2O3-H2O体系中由晶种水热生长的ZSM-5单晶形貌和线性生长速度. 晶体外部晶面为轴面(100)、(010)及坡面(101)、(011). 在互相贯穿连生的各单晶之间, 其对称轴互成倾角. 在160-220℃范围内, 晶体线性生长速度保持恒定. 观察到单晶生长的诱导期. 由线性生长速度与温度的关系计算了生长活化能. 研究了反应物组成对晶面法线生长速度的影响、母液成分变化, 并分析了晶体的化学成分. 按晶体生长的基本原理及ZSM-5结构, 构想了实际上可能存在的ZSM-5单晶的面结构, 并估算了不同晶面上生长位的密度及其分布. 从面结构及沸石生长基元角度讨论了单晶生长机理的DEA的作用, 支持了生长基元为高聚态硅(铝)阴离子团的假设.  相似文献   

10.
从结晶化学角度出发,研究了硫氰酸汞镉(简写为CMTC〕结构是的基本结构单无即H~gS~4和CdN~4R的结晶方位与晶体各簇晶面间的对应关系。在对CMTC的生长溶液结构测定和推理基础上,提出该晶体生长基元的结构形式和生长基元向各簇晶面上堆积的规律。进而讨论了该晶体生长习性的形成机理。实验表明,CMTC的生长溶液中存在着与晶体结构是相同的基团即阴离子多面体生长基元。随着生长条件和变化,生长基元的维度有所不同,不同维度的生长基元往各簇晶面上堆积速率也发生相应改变,从则解释了CMTC的生长习性和晶体形貌的多变性。  相似文献   

11.
The nucleation rate and propagation rate of steps on the {100} faces of polyethylene crystals have been determined. For single crystals, under conditions where the width of the {100} sectors remains constant during growth, it is confirmed that the growth is in regime I or the crossover region between regime I and II. In {110} twinned crystals, the {100} sectors are well developed and the width increases linearly with time; therefore, the growth in the twins must be in regime II. It is shown that the differing growth regimes of {100} faces in single crystals and twins allow the independent determination of the nucleation rate and the propagation rate of steps. The nucleation rate and propagation rate of steps on the {100} faces were determined from measurements of the constant width of the {100} faces in single crystals and the growth rate of the {100} faces in single crystals and twins. The observed rates show abnormal dependence on supercooling and concentration. The results are attributed to a weaker dependence of the constant width of {100} sectors on supercooling and concentration than predicted.  相似文献   

12.
Scanning confocal interference microscopy (SCIM) and molecular dynamics (MD) simulations were used to investigate the adsorption of the synthetic polypeptide poly(l-glutamic acid) (poly-glu) to calcium oxalate monohydrate (COM) crystals and its effect on COM formation. At low concentrations (1 μg/mL), poly-glu inhibits growth most effectively in ?001? directions, indicating strong interactions of the polypeptide with {121} crystal faces. Growth in ?010? directions was inhibited only marginally by 1 μg/mL poly-glu, while growth in ?100? directions did not appear to be affected. This suggests that, at low concentrations, poly-glu inhibits lattice-ion addition to the faces of COM in the order {121} > {010} ≥ {100}. At high concentrations (6 μg/mL), poly-glu resulted in the formation of dumbbell-shaped crystals featuring concave troughs on the {100} faces. The effects on crystal growth indicate that, at high concentrations, poly-glu interacts with the faces of COM in the order {100} > {121} > {010}. This mirrors MD simulations, which predicted that poly-glu will adsorb to a {100} terrace plane (most calcium-rich) in preference to a {121} (oblique) riser plane but will adsorb to {121} riser plane in preference to an {010} terrace plane (least calcium-rich). The effects of different poly-glu concentration on COM growth (1-6 μg/mL) may be due to variations between the faces in terms of growth mechanism and/or (nano)roughness, which can affect surface energy. In addition, 1 μg/mL might not be adequate to reach the critical concentration for poly-glu to significantly pin step movement on {100} and {010} faces. Understanding the mechanisms involved in these processes is essential for the development of agents to reduce recurrence of kidney stone disease.  相似文献   

13.
The rates of growth of polyethylene single crystals grown from dilute solution in hexadecane and tetradecanol have been measured over the temperature range Tc = 98–120°C by following the change in turbidity during crystallization of a suspension of crystals of known shape and final size. The rates decrease similarly with Tc in each solvent, but for a given supercooling crystals grow much faster in tetradecanol where the corresponding crystallization temperature is higher. Similarly, the rates are much higher in hexadecane than those previously reported from xylene at equivalent supercoolings but lower Tc. Changes in the corresponding crystal morphologies as Tc is raised are quantified in terms of the axial ratio and the degree of curvature of the nominally {100} faces, both of which increase with Tc. The results can be interpreted as showing a transition from regime I to regime II growth in both solvents, which agrees both qualitatively and quantitatively with the predictions of the nucleation-based kinetic theories. Such a transition has never before been reported for solution crystallization. Using this analysis, reasonable values are obtained for the crystal side-surface energy σ of 7.4–7.5 erg cm?2 and for the regime I substrate length L of 0.14 μm. No correlation is found between crystal morphology and growth rate and there are no discontinuous changes in morphology at the proposed transition points. The occurrence of curved crystal edges raises the fundamental issue of how to reconcile noncrystallographic growth surfaces with nucleation-controlled growth. A new approach to polymer crystal growth based on equilibrium surface roughening, which does not require nucleation, is therefore very pertinent in this respect and this is discussed.  相似文献   

14.
Hexagonal bipyramidal ruby (Al2O3:Cr) crystals were easily grown by the evaporation of MoO3 flux isothermally. The crystal growth was conducted by heating a mixture of solute (Al2O3 + 0.5 mass% Cr2O3) and flux (MoO3) at 1100 degrees C, followed by holding the solution at this temperature for 5 h. The ruby crystals obtained lengths up to 1.8 mm and widths up to 1.7 mm and were transparent-red. Their form was a bipyramid bounded by well-developed {113} faces. Molybdenum trioxide was found to be a very suitable flux for growing bipyramidal ruby crystals.  相似文献   

15.
Crystal growth of the metal–organic framework MOF‐5 was studied by atomic force microscopy (AFM) for the first time. Growth under low supersaturation conditions was found to occur by a two‐dimensional or spiral crystal growth mechanism. Observation of developing nuclei during the former reveals growth occurs through a process of nucleation and spreading of metastable and stable sub‐layers revealing that MOFs may be considered as dense phase structures in terms of crystal growth, even though they contain sub‐layers consisting of ordered framework and disordered non‐framework components. These results also support the notion this may be a general mechanism of surface crystal growth at low supersaturation applicable to crystalline nanoporous materials. The crystal growth mechanism at the atomistic level was also seen to vary as a function of the growth solution Zn/H2bdc ratio producing square terraces with steps parallel to the <100> direction or rhombus‐shaped terraces with steps parallel to the <110> direction when the Zn/H2bdc ratio was >1 or about 1, respectively. The change in relative growth rates can be explained in terms of changes in the solution species concentrations and their influence on growth at different terrace growth sites. These results were successfully applied to the growth of as‐synthesized cube‐shaped crystals to increase expression of the {111} faces and to grow octahedral crystals of suitable quality to image using AFM. This modulator‐free route to control the crystal morphology of MOF‐5 crystals should be applicable to a wide variety of MOFs to achieve the desired morphological control for performance enhancement in applications.  相似文献   

16.
Using the technique of extraction, single crystals have been obtained from polyethylene fractions isothermally crystallized from the melt at atmospheric pressure. It has been found that the lateral habit of single crystals changes in the vicinity of the transition temperature of growth regime (regime I–II): lenticular shape elongated in the direction of theb axis (type A) in the range of regime I and truncated lozenge with curved edges of {200} and {110} growth faces (type B) in that of regime II. The transition of lateral habit causes a drastic change in the width of {110} growth faces; {110} growth faces are well developed in type B crystals while they cannot be observed and must be very small in type-A crystals. It has been shown that the growth regime of the small {110} growth face of type-A crystals must be in regime I; hence the regime I–II transition can be explained as the result of this change in lateral habit (width of the {110} growth face).  相似文献   

17.
The morphology of micrometer-sized beta-hematin crystals (synthetic malaria pigment) was determined by TEM images and diffraction, and by grazing incidence synchrotron X-ray diffraction at the air-water interface. The needle-like crystals are bounded by sharp {100} and {010} side faces, and capped by {011} and, to a lesser extent, by {001} end faces, in agreement with hemozoin (malaria pigment) crystals. The beta-hematin crystals grown in the presence of 10% chloroquine or quinine took appreciably longer to precipitate and tended to be symmetrically tapered toward both ends of the needle, due to stereoselective additive binding to {001} or {011} ledges. Evidence, but marginal, is presented that additives reduce crystal mosaic domain size along the needle axis, based on X-ray powder diffraction data. Coherent grazing exit X-ray diffraction suggests that the mosaic domains are smaller and less structurally stable than in pure crystals. IR-ATR and Raman spectra indicate molecular based differences due to a modification of surface and bulk propionic acid groups, following additive binding and a molecular rearrangement in the environment of the bulk sites poisoned by occluded quinoline. These results provided incentive to examine computationally whether hemozoin may be a target of antimalarial drugs diethylamino-alkoxyxanthones and artemisinin. A variation in activity of the former as a function of the alkoxy chain length is correlated with computed binding energy to {001} and {011} faces of beta-hematin. A model is proposed for artemisinin activity involving hemozoin nucleation inhibition via artemisinin-beta-hematin adducts bound to the principal crystal faces. Regarding nucleation of hemozoin inside the digestive vacuole of the malaria parasite, nucleation via the vacuole's membranous surface is proposed, based on a reported hemozoin alignment. As a test, a dibehenoyl-phosphatidylcholine monolayer transferred onto OTS-Si wafer nucleated far more beta-hematin crystals, albeit randomly oriented, than a reference OTS-Si.  相似文献   

18.
Hierarchical TiO(2) nanospheres with controlled surface morphologies and dominant {001} facets were directly synthesized from Ti powder by a facile, one-pot, hydrothermal method. The obtained hierarchical TiO(2) nanospheres have a uniform size of 400-500?nm and remarkable 78?% fraction of {001} facets. The influence of the reaction temperature, amount of HF, and reaction time on the morphology and the exposed facets was systematically studied. A possible growth mechanism speculates that Ti powder first dissolves in HF solution, and then flowerlike TiO(2) nanostructures are formed by assembly of TiO(2) nanocrystals. Because of the high concentration of HF in the early stage, these TiO(2) nanostructures were etched, and hollow structures formed on the surface. After the F(-) ions were effectively absorbed on the crystal surfaces, {001} facets appear and grow steadily. At the same time, the {101} facets also grow and meet the {101} facets from adjacent truncated tetragonal pyramids, causing coalescence of these facets and formation of nanospheres with dominant {001} facets. With further extension of the reaction time, single-crystal {001} facets of hierarchical TiO(2) nanospheres are dissolved and TiO(2) nanospheres with dominant {101} facets are obtained. The photocatalytic activities of the hierarchical TiO(2) nanospheres were evaluated and found to be closely related to the exposed {001} facets. Owing to the special hierarchical architecture and high percentage of exposed {001} facets, the TiO(2) nanospheres exhibit much enhanced photocatalytic efficiency (almost fourfold) compared to P25 TiO(2) as a benchmark material. This study provides new insight into crystal-facet engineering of anatase TiO(2) nanostructures with high percentage of {001} facets as well as opportunities for controllable synthesis of 3D hierarchical nanostructures.  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