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 31 毫秒
1.
A modified adaptive immune optimization algorithm (AIOA) is designed for optimization of Cu–Au and Ag–Au bimetallic clusters with Gupta potential. Compared with homoatom clusters, there are homotopic isomers in bimetallic cluster, so atom exchange operation is presented in the modified AIOA. The efficiency of the algorithm is tested by optimization of CunAu38‐n (0 ≤ n ≤ 38). Results show that all the structures with the putative global minimal energies are successfully located. In the optimization of AgnAu55‐n (0 ≤ n ≤ 55) bimetallic clusters, all the structures with the reported minimal energies are obtained, and 36 structures with even lower potential energies are found. On the other hand, with the optimized structures of CunAu55‐n, it is shown that all 55‐atom Cu–Au bimetallic clusters are Mackay icosahedra except for Au55, which is a face‐centered cubic (fcc)‐like structure; Cu55, Cu12Au43, and Cu1Au54 have two‐shell Mackay icosahedral geometries with Ih point group symmetry. © 2009 Wiley Periodicals, Inc. J Comput Chem 2009  相似文献   

2.
The solid solution CeRu1–xPdxAl was synthesized for x = 0.1–0.9 from the elements by arc‐melting and subsequent annealing and characterized by powder X‐ray diffraction. All members crystallize in the orthorhombic LaNiAl type structure, space group Pnma. The lattice parameters range from a = 718–722, b = 412–426, and c = 1588–1620 pm, but no linear change of the lattice parameters was found. Magnetic measurements reveal intermediate cerium valences, which change to more trivalent cerium with increasing Pd content. The susceptibility data was interpreted by either the Inter‐Configuration Fluctuation (ICF) model or the Curie‐Weiss law.  相似文献   

3.
A second type of cation (Mg2+, Ca2+) was introduced into BaF2 by low‐temperature atomic beam deposition. The structure evolution from low‐temperature (–150 °C) amorphous deposits to high‐temperature (< 1000 °C) annealed crystalline phases was studied by in‐situ transmission electron microscopy and X‐ray diffraction. Amorphous (Ba0.5, Ca0.5)F2 crystallizes in a first step to metastable solid solution phase (fluorite‐type), which then decomposes into the pure phases of CaF2 and BaF2 at higher temperature. The crystallization behavior of amorphous (BaxMg1–x)F2 is completely different. When the Mg/Ba atomic ratio is around 1:1, the mixture transforms to the ternary compound BaMgF4 at annealing, and no decomposition occurs by further heating up to 1000 °C. When the Ba concentration is below 15 % in atomic ratio (x < 0.15), the mixture forms a solid solution phase (rutile type) with the lattice expanded by +1 % compared to rutile type MgF2. The difference between the phase evolutions of the two mixture systems is discussed.  相似文献   

4.
[PtIn6][GaO4]2 – The First Oxide Containing [PtIn6] Octahedra. Preparation, Characterisation, and Rietveld Refinement – With a Remark to the Solid Solution Series [PtIn6][GaO4]2‐x[InO4]x (0 < x ≤ 1) The novel oxides [PtIn6][GaO4]2–x[InO4]x (0 < x ≤ 1) are formed by heating intimate mixtures of Pt, In, In2O3, and Ga2O3 in the corresponding stoichiometric ratio in corundum crucibles under an atmosphere of argon (1220 K, 70 h). The compounds are black, stable in air at room temperature, reveal a semiconducting behaviour, and decompose only in oxidizing acids. X‐ray powder diffraction patterns can be indexed by assuming a face centered cubic unit cell with lattice parameters ranging from a = 1001.3(1) pm (x = 0) to a = 1009.3(1) pm (x = 1). According to a Rietveld refinement [PtIn6][GaO4]2 crystallizes isotypic to the mineral Pentlandite (Fm3m, Z = 4, R(profile) = 6.11%, R(intensity) = 3.95%). The characteristic building units are isolated [PtIn6]10+ octahedra which are linked via [GaO4]5– tetrahedra to a three dimensional framework. Starting from [PtIn6][GaO4]2 the substitution of Ga3+ ions by larger In3+ ions leads to the formation of a solid solution series according to the general formula [PtIn6][GaO4]2–x[InO4]x and becomes apparent in an increase of the lattice parameter.  相似文献   

5.
The synthesis and structure of atomically precise Au130?xAgx (average x=98) alloy nanoclusters protected by 55 ligands of 4‐tert‐butylbenzenethiolate are reported. This large alloy structure has a decahedral M54 (M=Au/Ag) core. The Au atoms are localized in the truncated Marks decahedron. In the core, a drum of Ag‐rich sites is found, which is enclosed by a Marks decahedral cage of Au‐rich sites. The surface is exclusively Ag?SR; X‐ray absorption fine structure analysis supports the absence of Au?S bonds. The optical absorption spectrum shows a strong peak at 523 nm, seemingly a plasmon peak, but fs spectroscopic analysis indicates its non‐plasmon nature. The non‐metallicity of the Au130?xAgx nanocluster has set up a benchmark to study the transition to metallic state in the size evolution of bimetallic nanoclusters. The localized Au/Ag binary architecture in such a large alloy nanocluster provides atomic‐level insights into the Au?Ag bonds in bimetallic nanoclusters.  相似文献   

6.
Abstract. The ternary Zintl phase Ca3Ag1+xGe3–x (x = 1/3) was synthesized by the high‐temperature solid‐state technique and its crystal structure was refined from single‐crystal diffraction data. The compound Ca3Ag1.32Ge2.68(1) adopts the Sc3NiSi3 type structure, crystal data: space group C2/m, a = 10.813(1) Å, b = 4.5346(4) Å, c = 14.3391(7) Å, β = 110.05(1)° and V = 660.48(10) Å3 for Z = 4. Its structure can be interpreted as an intergrowth of fragments cut from the CaGe (CrB‐type) and the CaAg1+xGe1–x (TiNiSi‐type) structures, and it therefore represents an alkaline‐earth member of the structure series with the general formula R2+nT2X2+n with n = 4. Unlike the rare‐earth homologues that are fully ordered phases, one seventh of the atomic sites in the unit cell of the title compound are mixed occupied (roughly 2/3Ge and 1/3Ag), and this can be explained by the Zintl concept. The alloying of this phase using aluminum metal yielded the isotypic solid solution Ca3(Ag/Al)1+xGe3–x, in which the aluminum for silver substitution is strictly localized in the TiNiSi substructure, revealing the very different functionality of the two building blocks.  相似文献   

7.
Reaction of AgNO3 and 2,2′‐bipyridine (bipy) under ultrasonic treatment gave the title compound, [Ag(C10H8N2)(NH3)]NO3. The crystal structure consists of dimers formed by two symmetry‐related AgI–bipy monomers connected through intra‐dimer π–π stacking and ligand‐unsupported Ag...Ag interactions. A crystallographic C2 axis passes through the mid‐point of and is perpendicular to the Ag...Agi(−x + 1, y, −z + ) axis. In addition, each AgI cation is coordinated by one chelating bipy ligand and one ammine ligand, giving a trigonal coordination environment capped by the symmetry‐equivalent Ag atom. Molecules are assembled by Ag...Ag, π–π, hydrogen‐bond (N—H...O and C—H...O) and weak Ag...π interactions into a three‐dimensional framework. Comparing the products synthesized under different mechanical treatments, we found that reaction conditions have a significant influence on the resulting structures. The luminescence properties of the title compound are also discussed.  相似文献   

8.
The crystal structures of MgAl2–xGaxO4 (0 ≤ x ≤ 2) spinel solid solutions (x = 0.00, 0.38, 0.76, 0.96, 1.52, 2.00) were refined using 27Al MAS NMR measurements and single crystal X‐ray diffraction technique. Site preferences of cations were investigated. The inversion parameter (i) of MgAl2O4 (i = 0.206) is slightly larger than given in previous studies. It is considered that the difference of inversion parameter is caused by not only the difference of heat treatment time but also some influence of melting with a flux. The distribution of Ga3+ is little affected by a change of the temperature from 1473 K to 973 K. The degree of order‐disorder of Mg2+ or Al3+ between the fourfold‐ and sixfold‐coordinated sites is almost constant against Ga3+ content (x) in the solid solution. A compositional variable of the Ga/(Mg + Ga) ratio in the sixfold‐coordinated site has a constant value through the whole compositional range: the ratio is not influenced by the occupancy of Al3+. The occupancy of Al3+ is independent of the occupancy of Ga3+, though it depends on the occupancy of Mg2+ according to thermal history. The local bond lengths were estimated from the refined data of solid solutions. The local bond length between specific cation and oxygen corresponds with that expected from the effective ionic radii except local Al–O bond length in the fourfold‐coordinated site and local Mg–O bond length in the sixfold‐coordinated site. The local Al–O bond length in the fourfold‐coordinated site (1.92 Å) is about 0.15 Å longer than the expected bond length. This difference is induced by a difference in site symmetry of the fourfold‐coordinated site. The nature that Al3+ in spinel structure occupies mainly the sixfold‐coordinated site arises from the character of Al3+ itself. The local Mg–O bond length in the sixfold‐coordinated site (2.03 Å) is about 0.07 Å shorter than the expected one. Difference Fourier synthesis for MgGa2O4 shows a residual electron density peak of about 0.17 e/Å3 in height on the center of (Ga0.59 Mg0.41)–O bond. This peak indicates the covalent bonding nature of Ga–O bond on the sixfold‐coordinated site in the spinel structure.  相似文献   

9.
Deep blue‐violet colored powder samples of Ag2ZnZr2F14 were synthesized by heating Zn(NO3)2·4H2O, Ag and ZrOCl2·8H2O at 300 °C under fluorine atmosphere. The crystal structure of Ag2ZnZr2F14 was refined from X‐ray powder diffraction data using the Rietveld method (C2/m, a = 9.0206(1) Å, b = 6.6373(1) Å, c = 9.0563(1) Å, β = 90.44(1)°, Z = 2). The structure is derived from the isotypic Ag3Zr2F14 by replacing only one of the two crystallographically different Ag2+ ions with Zn2+ ions, thus leading to discrete Ag2F7 dimers. These dimers are connected via nearly linear Ag–F···F–Ag bridges with short F···F distances of 2.33 Å to form two‐legged ladders. Magnetic susceptibility measurements and density functional calculations show that the two Ag2+ ions in each Ag2F7 dimer are strongly coupled antiferromagnetically.  相似文献   

10.
Precise atomic structure of metal nanoclusters (NCs) is fundamental for elucidating the structure–property relationships and the inherent size‐evolution principles. Reported here is the largest known FCC‐based (FCC=face centered cubic) silver nanocluster, [Ag100(SC6H33,4F2)48(PPh3)8]?: the first all‐octahedral symmetric nesting Ag nanocluster with a four‐layered Ag6@Ag38@Ag48S24@Ag8S24P8 structure, consistent symmetry elements, and a unique rhombicuboctahedral morphology distinct from theoretical predictions and previously reported FCC‐based Ag clusters. DFT studies revealed extensive interlayer interactions and degenerate frontier orbitals. The FCC‐based Russian nesting doll model constitutes a new platform for the study of the size‐evolution principles of Ag NCs.  相似文献   

11.
The structures, thermal and physical properties of ion conducting polymorphic Ag5Te2Cl1?xBrx and Ag5Te2?ySyCl have been investigated. A maximum substitution degree of x = 0.65 and y = 0.3 was derived from X‐ray powder diffraction. Mixtures of silver halides, silver chalcogenides and Ag3TeBr were observed for higher substitution degrees. Both silver chalcogenide halide systems show a Vegard type behaviour. Single crystal structure determinations of selected materials were performed at different temperatures to analyse the silver distribution in the tetragonal high temperature α‐ and the monoclinic room temperature β‐phases. After non‐harmonic refinement of the silver positions detailed joint probability density function analysis (jpdf) and determination of one particle potentials (opp) were carried out to investigate the diffusion pathways and bottlenecks of ion transport for those materials. A preferred anisotropic ion transport along the diffusion pathways for the α‐ and 1D zig‐zag diffusion pathways for the β‐phases were found. αβ and βγ phase transitions were determined by DSC and DTA methods and conductivities were measured using temperature dependent impedance spectroscopy. The substitution of tellurium by sulphur lowered the αβ phase transition from 334 K (Ag5Te2Cl) to 270 K (Ag5Te1.8S0.2Cl) while the opposite trend was found for the Ag5Te2Cl1?xBrx phases. The αβ phase transition of Ag5Te2Cl0.35Br0.65 at 343 K represents the highest transition observed for the silver chalcogenide halides under discussion. Total conductivities of approx. 1 Ω?1 cm?1 (α‐Ag5Te2Cl0.5Br0.5) and 0.24 Ω?1 cm?1 (α‐Ag5Te1.8S0.2Cl) at 473 K were found being slightly higher (Br) and lower (S) than the conductivity observed for α‐Ag5Te2Cl. A conductivity jump of more than two orders of magnitude, related to the αβ phase transitions, within the temperature range from 270 to 343 K is adjustable by simple variation of the composition and is therefore an extraordinary feature of these materials. The total conductivity is linearly correlated to the volume of the anion substructure and can be varied within more than half an order of magnitude.  相似文献   

12.
The boron‐centered mixed‐halide zirconium cluster phases were obtained from reactions of appropriate amounts of NaCl or ACl2 (with A = Ca, Sr, Ba), ZrCl4, ZrI4, Zr (powder), and elemental B in sealed tantalum containers at 800–850 °C. Single crystals of NaZr6Cl10.94(1)I3.06B and Sr0.5Zr6Cl11.34(2)I2.66B have been characterized by X‐ray diffraction at room temperature (cubic, Pa 3, Z = 4, a = 1331.3(1) pm, and a = 1326.2(2) pm, respectively). The Zr6 octahedra in these phases are centered by boron and three‐dimensionally connected by exo‐iodide atoms which bridge simultaneously three octahedra. Six out of the twelve inner chlorides (one symmetry independent site) which bridge the edges of each octahedron, but are not involved in inter‐cluster bridging (according to AZr6Ia–a–a6/2(Cl12–xIx)iB), can be completely substituted by iodide. Such a substitution is not possible for the remaining six inner halides on the second symmetry independent site, because the larger iodide atom on this site would experience strong repulsive forces from short anionic contacts. This gives the phase width of NaZr6Cl12–xI2+xB to x È 6. The size of the voids that cover the cations A limits this structure to members with A = Na, Ca, Sr, and Ba. This structure type requires the existence of two differently sized halide types simultaneously.  相似文献   

13.
In the course of a study on the role of magnesium in polar zincides of the heavier alkaline‐earth elements, three intermetallic phases of the ternary system Ca–Mg–Zn were synthesized from melts of the elements and their structures were determined by means of single‐crystal X‐ray data. Starting from the binary zincide CaZn11, the phase width of the BaCd11‐type structure reaches up to the fully ordered stoichiometric compound CaMgZn10 [tI48, space group I41/amd, a = 1082.66(6), c = 688.95(5) pm, Z = 4, R1 = 0.0239]. The new compound CaMgZn5 (oP28, space group Pnma, a = 867.48(3), b = 530.37(5), c = 1104.45(9) pm, Z = 4, R1 = 0.0385) crystallizes in the CeCu6‐type structure, exhibits no Mg/Zn phase width and has no binary border equivalent in the system Ca–Mg–Zn. Similar to the situation in CaMgZn10, one M position of the aristotype has a slightly larger coordination sphere (CN = 14) and is accordingly occupied by the larger Mg atoms. The third phase, Ca2+xMg6–xyZn15+y (hP92, space group P63/mmc, a = 1476.00(5), c = 881.01(4) pm, Z = 4, R1 = 0.0399 for Ca2.67Mg5.18Zn15.15) forms the hexagonal Sm3Mg13Zn30‐type structure also known as μ‐MgZnRE or S phase. A small phase width (x = 0–0.67; y = 0–0.58) is due to the slightly variable Ca or Zn content of the two Mg positions. The structure is described as an intergrowth of the hexagonal MgZn2 Laves phase and the CaZn2 structure (KHg2‐type). All compounds exhibit strong Zn–Zn and polar Mg–Zn covalent bonds, which are visible in the calculated electron density maps. Their structures are thus herein described using the full space tilings of [Zn4] and [MgZn3] tetrahedra, which are fused to polyanions consisting of tetrahedra stars, icosahedra segments etc. and the large (CN = 18–22) Ca cation coordination polyhedra. Pseudo bandgaps apparent in the tDOS are compatible with the narrow v.e./M ranges observed for other isotypic members of the three structure types.  相似文献   

14.
The rod‐shaped Au25 nanocluster possesses a low photoluminescence quantum yield (QY=0.1 %) and hence is not of practical use in bioimaging and related applications. Herein, we show that substituting silver atoms for gold in the 25‐atom matrix can drastically enhance the photoluminescence. The obtained AgxAu25?x (x=1–13) nanoclusters exhibit high quantum yield (QY=40.1 %), which is in striking contrast with the normally weakly luminescent AgxAu25?x species (x=1–12, QY=0.21 %). X‐ray crystallography further determines the substitution sites of Ag atoms in the AgxAu25?x cluster through partial occupancy analysis, which provides further insight into the mechanism of photoluminescence enhancement.  相似文献   

15.
Preparation, Crystal Structure, and Temperature Dependence of the Homogeneity Range of the Phase Ca2-xAs1-xBr1+x and Thermal Behaviour of Ca3AsBr3 The phase (Ca2-xx)(As1-xBr1+x) (yellow, NaCl type lattice, x ? degree of substitution) was prepared from “Ca3As2” and CaBr2 in different molar ratios in steel ampoules under argon at 900 and 950°C resp. The lattice constant as a function of the composition, the homogeneity range, and dependence of the bromine rich phase boundary on the temperature were determined. The structure was deduced from single crystal X-ray investigation and density measurements at different compositions. The thermal behaviour of Ca3AsBr3 (colourless, isotypic to Mg3NF3, prepared at 850°C) was studied by annealing samples in molybdenum ampoules under argon in the temperature range 900–1250°C and by differential thermal analysis. From the experimental results a section of the phase diagram Ca3As2?CaBr2 was constructed.  相似文献   

16.
A series of Mg‐Zr mixed oxides with different nominal Mg/ (Mg+Zr) atomic ratios, namely 0, 0.1, 0.2, 0.4, 0.85, and 1, is prepared by alcogel methodology and fundamental insights into the phases obtained and resulting active sites are studied. Characterization is performed by X‐ray diffraction, transmission electron microscopy, X‐ray photoelectron spectroscopy, N2 adsorption–desorption isotherms, and thermal and chemical analysis. Cubic MgxZr1?xO2?x solid solution, which results from the dissolution of Mg2+ cations within the cubic ZrO2 structure, is the main phase detected for the solids with theoretical Mg/ (Mg+Zr) atomic ratio ≤0.4. In contrast, the cubic periclase (c‐MgO) phase derived from hydroxynitrates or hydroxy precursors predominates in the solid with Mg/(Mg+Zr)=0.85. c‐MgO is also incipiently detected in samples with Mg/(Mg+Zr)=0.2 and 0.4, but in these solids the c‐MgO phase mostly arises from the segregation of Mg atoms out of the alcogel‐derived c‐MgxZr1?xO2?x phase during the calcination process, and therefore the species c‐MgO and c‐MgxZr1?xO2?x are in close contact. Regarding the intrinsic activity in furfural–acetone aldol condensation in the aqueous phase, these Mg? O? Zr sites located at the interface between c‐MgxZr1?xO2?x and segregated c‐MgO display a much larger intrinsic activity than the other noninterface sites that are present in these catalysts: Mg? O? Mg sites on c‐MgO and Mg? O? Zr sites on c‐MgxZr1?xO2?x. The very active Mg? O? Zr sites rapidly deactivate in the furfural–acetone condensation due to the leaching of active phases, deposition of heavy hydrocarbonaceous compounds, and hydration of the c‐MgO phase. Nonetheless, these Mg‐Zr materials with very high specific surface areas would be suitable solid catalysts for other relevant reactions catalyzed by strong basic sites in nonaqueous environments.  相似文献   

17.
The new ternary phase Eu2–xMg2–yGe3 (x = 0.1, y = 0.5) was obtained by solid‐state synthesis and the structure determined by means of Single Crystal X‐ray Diffraction. The compound crystallizes with the orthorhombic space group Cmcm (no. 63) having structural features of the low‐temperature modification of LaSi. The crystal structure contains two different types of germanium anions: isolated Ge4– and $\rm^{2}_{\infty}$ [Ge2–xy] chains. The cation substructure is partially disordered and is best represented assuming a split position. The chemical bonding is well represented by the Zintl‐Klemm concept. Resistivity measurements reveal that the compound is metallic. DFT band structure calculations were carried out on the ideal stoichiometric compound Eu2Mg2Ge3, showing that this model (x = 0; y = 0) would be also metallic as a consequence of the ecliptic stacking of anions. Susceptibility and specific heat measurements evidence the presence of weak, and probably frustrated, antiferromagnetic interactions between disordered europium atoms.  相似文献   

18.
The intermetallic phase SrNi2–xSb2 was synthesized by arc melting mixtures of the elements and subsequent annealing under argon, and its structure was investigated by means of both powder and single‐crystal X‐ray diffraction methods. It crystallizes with the ThCr2Si2 type (tI10, I4/mmm) and has a homogeneity range of x = 0.15(1)–0.28(1), which does not include the exact stoichiometric 1:2:2 composition. The crystal structure of the phase SrNi2–xSb2 is built from layers of edge‐sharing Ni1–xSb4 tetrahedra, which are separated by Sr atoms along the c direction. Magnetic measurements of SrNi2–xSb2 showed no superconductive transition above 1.8 K.  相似文献   

19.
Novel ternary phases, (Pd1?xZnx)18(Zn1?yAly)86?δ (0≤x≤0.162, 0.056≤y≤0.088, 0≤δ≤4), which adopt a superstructure of the γ‐brass type (called γ′‐brass), have been synthesized from the elements at 1120 K. Single‐crystal X‐ray structural analysis reveals a phase width (F$\bar 4$ 3m, a=18.0700(3)–18.1600(2) Å, Pearson symbols cF400–cF416), which is associated with structural disorder based on both vacancies as well as mixed site occupancies. These structures are constructed of four independent 26‐atom γ‐clusters per primitive unit cells and centered at the four special positions A (0, 0, 0), B (1/4, 1/4, 1/4), C (1/2, 1/2, 1/2) and D (3/4, 3/4, 3/4). Two of these, centered at B and C , are completely ordered Pd4Zn22 clusters, whereas the other two, centered at A and D , contain all structural disorder in the system. According to our single‐crystal X‐ray results, Al substitutions are restricted to the A ‐ and D ‐centered clusters. Moreover, the outer tetrahedron (OT) site of the 26‐atom cluster at D is completely vacant at the Al‐rich boundary of these phases. Electronic structure calculations, using the tight‐binding linear muffin‐tin orbital atomic‐spheres approximation (TB‐LMTO‐ASA) method, on models of these new, ternary γ′‐brass phases indicate that the observed chemical compositions and atomic distributions lead to the presence of a pseudogap at the Fermi level in the electronic density of states curves, which is consistent with the Hume‐Rothery interpretation of γ‐brasses, in general.  相似文献   

20.
Tetrasilver telluride sulfate was obtained as a black air‐stable polycrystalline powder; its structure was determined from X‐ray powder diffraction data. The new compound crystallizes in the cubic space group P213, with the unit cell parameter a = 8.6263(2) Å and Z = 4. The crystal structure of Ag4Te(SO4) is composed of a positively charged silver‐telluride three‐dimensional framework, in which the isolated tetrahedral SO42– anions are embedded. The framework features an irregular coordination for tellurium atoms with a coordination number of six and manifold Ag–Ag contacts ranging from 2.99 to 3.14 Å. The distortion of the SO42– anion as well as the interactions within the framework and between the framework and the SO42– anions are analyzed with the help of the structural data, vibrational spectra, and band structure calculations.  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