共查询到20条相似文献,搜索用时 31 毫秒
1.
《Journal of solid state chemistry》1986,62(2):133-137
Four ternary phases MPtSi (M = Ca, Eu, Sr, Ba) have been shown to crystallize in the LaIrSi-type structure (space group P213). This ternary structure is a derivative structure of the binary SrSi2-type structure (space group P4332 or P4132). In the MPtSi series the LaIrSi-type structure has a stability range for metals with radii from rCa = 1.973 Å to rBa = 2.243 Å in contrast to MSi2 compounds which exist with the SrSi2-type structure only from rSr = 2.151Å to rBa 2.243 Å. From a single-crystal investigation on CaPtSi remarkably short PtSi distances of 2.30 Å (3x) are obtained. Structural relations are discussed. 相似文献
2.
Dr. Glauco S. Maciel Dr. Whualkuer Lozano Bartra Dr. Yutao Xing Dr. Nikifor Rakov 《Chemphyschem》2022,23(2):e202100517
There is a large interest in luminescent materials for application as temperature sensors. In this scenario, we investigate the performance of neodymium-doped alkaline-earth fluoride (Nd3+:MF2; M=Ba, Ca, Sr) crystalline powders prepared by combustion synthesis for optical temperature-sensing applications based on the luminescence intensity ratio (LIR) technique. We observe that the near-infrared luminescence spectral profile of Nd3+ changes with the temperature in a way that its behavior is suitable for optical thermometry operation within the first biological window. We also observe that the thermometric sensitivities of all studied samples change depending on the spectral integration range used in the LIR analysis. Nd3+:CaF2 presents the largest sensitivity values, with a maximum absolute sensitivity of 6.5×10−3/K at 824 K and a relative sensitivity of 1.71 %/K at human-body temperature (310 K). The performance of CaF2 for optical thermometry is superior to that of β-NaYF4, a standard material commonly used for optical bioimaging and temperature sensing, and on par with the most efficient oxide nanostructured materials. The use of thermometry data to help understand structural properties via Judd-Ofelt intensity standard parameters is also discussed. 相似文献
3.
我们合成了六种Eu2+激活的碱土金属氟卤化物MFX:Eu2+(M=Ca、Sr或Ba;X=Cl、Br或I)。研究了它们的荧光发射光谱和激发光谱,讨论了Eu2+离子的跃迁发射随基质晶体组成和结构变化的规律。根据晶体场理论,按照C4v点对称性,计算得到在MFCl:Eu2+(M=Ca、Sr或Ba)晶体中Eu2+离子的4?65d1激发态能级分裂的数值。 相似文献
4.
Two novel, noncentrosymmetric borate fluorides, Sr(3)B(6)O(11)F(2) and Ba(3)B(6)O(11)F(2), have been synthesized hydrothermally and their structures determined. The compounds are isostructural, crystallizing in space group P2(1), having lattice parameters of a = 6.4093 (13) ?, b = 8.2898 (17) ?, c = 9.3656 (19) ?, and β = 101.51 (3)° for Sr(3)B(6)O(11)F(2) and a = 6.5572 (13) ?, b = 8.5107 (17) ?, c = 9.6726 (19) ?, and β = 101.21 (3)° for Ba(3)B(6)O(11)F(2). The structure consists of a complex triple-ring borate framework having aligned triangular [BO(3)] groups that impart polarity. Fluorine atoms are bound only to the alkaline-earth metals and are not part of the borate framework, resulting in a vastly different structure from those of the hydrated borates Sr(3)B(6)O(11)(OH)(2) and Ba(3)B(6)O(11)(OH)(2) with similar formulas. The title compounds are transparent to nearly 200 nm, making them potentially useful for deep-ultraviolet nonlinear-optical applications. 相似文献
5.
After high-pressure/high-temperature treatment (40 kbar at 1000–1500°C) and quenching to ambient conditions CaSi2, EuSi2, and SrSi2 crystallize in the α-ThSi2 type of structure. Lattice constants and positional parameters have been determined by X-ray powder technique. Structural relations are discussed. 相似文献
6.
7.
Gadzhiev S. M. Shabanov O. M. Magomedova A. O. Dzhamalova S. A. 《Russian Journal of Electrochemistry》2003,39(4):386-390
The electroconductivity of molten mixtures of calcium, strontium, and barium chlorides with potassium chloride (component concentrations 0, 25, 50, 75, 100 mol %) is studied as a function of the electric field strength. Isotherms of extreme high-voltage conductivities of the mixtures are an additive function of the composition, as opposed to isotherms of low-voltage conductivity, which exhibit considerable deviations and pass through minimums. 相似文献
8.
Sebastian Wendl Lucien Eisenburger Dr. Philipp Strobel Daniel Günther Dr. Jonathan P. Wright Dr. Peter J. Schmidt Prof. Dr. Oliver Oeckler Prof. Dr. Wolfgang Schnick 《Chemistry (Weinheim an der Bergstrasse, Germany)》2020,26(32):7292-7298
The nitridophosphates AEP8N14 (AE=Ca, Sr, Ba) were synthesized at 4–5 GPa and 1050–1150 °C applying a 1000 t press with multianvil apparatus, following the azide route. The crystal structures of CaP8N14 and SrP8N14 are isotypic. The space group Cmcm was confirmed by powder X-ray diffraction. The structure of BaP8N14 (space group Amm2) was elucidated by a combination of transmission electron microscopy and diffraction of microfocused synchrotron radiation. Phase purity was confirmed by Rietveld refinement. IR spectra are consistent with the structure models and the chemical compositions were confirmed by X-ray spectroscopy. Luminescence properties of Eu2+-doped samples were investigated upon excitation with UV to blue light. CaP8N14 (λem=470 nm; fwhm=1380 cm−1) and SrP8N14 (λem=440 nm; fwhm=1350 cm−1) can be classified as the first ultra-narrow-band blue-emitting Eu2+-doped nitridophosphates. BaP8N14 shows a notably broader blue emission (λem=417/457 nm; fwhm=2075/3550 cm−1). 相似文献
9.
在空气中采用高温固相反应方法合成的17MO-(8-x-y)-75B2O3-xGd2O3(MLBEG,M-Mg,Ca,Sr,Ba)玻璃,在紫外光(λex=350nm)激发下发射蓝光和红光,在绿色光(λex=532nm)激发下发射红光,电子自旋共振谱研究表明玻璃体系中有Eu^2 离子存在,蓝色区的宽带发射是Eu^2 离子的5d-4f跃迁发射:红色区的窄带发射是Eu^3 离子的5Do-7FJ(J=1,2,3,4)跃迁发射,发现玻璃中的碱土金属离子对Eu^3 /Eu^2 离子的比例有很大影响,选择不同的碱土金属离子可以调节玻璃蓝色光和红色光的相对发射强度,MLBEG玻璃的发光性质可用于转换太阳能,增强植物的光合作用。 相似文献
10.
Frank Kubel Nicole Wandl Mariana Pantazi Vincenza D'Anna Hans Hagemann 《无机化学与普通化学杂志》2013,639(6):892-898
The crystal structures of the M2NaIO6 series (M = Ca, Sr, Ba), prepared at 650 °C by ceramic methods, were determined from conventional laboratory X‐ray powder diffraction data. Synthesis and crystal growth were made by oxidizing I– with O2(air) to I7+ followed by crystal growth in the presence of NaF as mineralizator, or by the reaction of the alkali‐metal periodate with the alkaline‐earth metal hydroxide. All three compounds are insoluble and stable in water. The barium compound crystallizes in the cubic space group Fm3m (no. 225) with lattice parameters of a = 8.3384(1) Å, whereas the strontium and calcium compounds crystallize in the monoclinic space group P21/c (no. 14) with a = 5.7600(1) Å, b = 5.7759(1) Å, c = 9.9742(1) Å, β = 125.362(1)° and a = 5.5376(1) Å, b = 5.7911(1) Å, c = 9.6055(1) Å, β = 124.300(1)°, respectively. The crystal structure consists of either symmetric (for Ba) or distorted (for Sr and Ca) perovskite superstructures. Ba2NaIO6 contains the first perfectly octahedral [IO6]5– unit reported. The compounds of the ortho‐periodates are stable up to 800 °C. Spectroscopic measurements as well as DFT calculations show a reasonable agreement between calculated and observed IR‐ and Raman‐active vibrations. 相似文献
11.
采用高温固相法制备了Eu2+/Mn2+单激活和共激活的M3MgSi2O8-M2SiO4(M=Ba,Ca)两相荧光粉.通过X射线衍射(XRD)和荧光光谱(PL)对样品材料的晶体结构和光谱性能进行了表征.XRD测试结果表明所合成的样品具有M3MgSi2O8和M2SiO4两种晶相结构.PL测试显示,Eu2+在Ba3MgSi2O8-Ba2SiO4体系中发射442和502nm两个波带的光;而Eu2+在Ca2+部分取代Ba2+的BaCa2MgSi2O8-Ba1.31Ca0.69SiO4体系中发射420~520nm的连续波带,并且激发光谱向长波扩展,更加适用于被InGaN芯片(395 nm)激发.通过改变Mn2+的掺杂量可制得颜色可调的BaCa2MgSi2O8-Ba1.31Ca0.69SiO4:Eu2+,Mn2+白光荧光粉. 相似文献
12.
采用溶胶-凝胶法制备碱土金属钛酸盐MTiO3(M=Mg,Ca,Sr,Ba),并进一步与TiO2固相法复合制备MTiO3-TiO2异质结型复合光催化剂.以光催化降解亚甲基蓝(MB)为探针,评价了MTiO3和MTiO3-TiO2光催化剂的活性变化.结果表明,紫外光条件下碱土金属钛酸盐MTiO3的光催化活性顺序为:CaTiO3>BaTiO3>SrTiO3>MgTiO3,钙钛矿化合物的容忍因子、电负性以及催化剂的吸附性能都影响催化剂的降解效率.MTiO3与TiO2复合后形成的异质结复合光催化剂的催化活性得到显著的提高,催化剂浓度1.0g/L时,光催化反应1h后,MB(25mg/L)的降解率分别为82.6%,99.8%,93.7%,97.3%,异质结复合光催化剂活性顺序与MTiO3一致.光催化活性的提高与异质结界面形成电荷定向流动,促进光生电子、空穴的分离有关. 相似文献
13.
Chen Chen Meng-hui Wang Lin-Yan Feng Lian-Qing Zhao Jin-Chang Guo Hua-Jin Zhai Zhong-hua Cui Sudip Pan Gabriel Merino 《Chemical science》2022,13(27):8045
The occurrence of planar hexacoordination is very rare in main group elements. We report here a class of clusters containing a planar hexacoordinate silicon (phSi) atom with the formula SiSb3M3+ (M = Ca, Sr, Ba), which have D3h (1A1′) symmetry in their global minimum structure. The unique ability of heavier alkaline-earth atoms to use their vacant d atomic orbitals in bonding effectively stabilizes the peripheral ring and is responsible for covalent interaction with the Si center. Although the interaction between Si and Sb is significantly stronger than the Si–M one, sizable stabilization energies (−27.4 to −35.4 kcal mol−1) also originated from the combined electrostatic and covalent attraction between Si and M centers. The lighter homologues, SiE3M3+ (E = N, P, As; M = Ca, Sr, Ba) clusters, also possess similar D3h symmetric structures as the global minima. However, the repulsive electrostatic interaction between Si and M dominates over covalent attraction making the Si–M contacts repulsive in nature. Most interestingly, the planarity of the phSi core and the attractive nature of all the six contacts of phSi are maintained in N-heterocyclic carbene (NHC) and benzene (Bz) bound SiSb3M3(NHC)6+ and SiSb3M3(Bz)6+ (M = Ca, Sr, Ba) complexes. Therefore, bare and ligand-protected SiSb3M3+ clusters are suitable candidates for gas-phase detection and large-scale synthesis, respectively.The global minimum of SiSb3M3+ (M = Ca, Sr, Ba) is a D3h symmetric structure containing an elusive planar hexacoordinate silicon (phSi) atom. Most importantly, the phSi core remains intact in ligand protected environment as well.Exploring the bonding capacity of main-group elements (such as carbon or silicon) beyond the traditional tetrahedral concept has been a fascinating subject in chemistry for five decades. The 1970 pioneering work of Hoffmann and coworkers1 initiated the field of planar tetracoordinate carbons (ptCs), or more generally, planar hypercoordinate carbons. The past 50 years have witnessed the design and characterization of an array of ptC and planar pentacoordinate carbon (ppC) species.2–14 However, it turned out to be rather challenging to go beyond ptC and ppC systems. The celebrated CB62− cluster and relevant species15,16 were merely model systems because C avoids planar hypercoordination in such systems.17,18 In 2012, the first genuine global minimum D3h CO3Li3+ cluster was reported to have six interactions with carbon in planar form, although electrostatic repulsion between positively charged phC and Li centers and the absence of any significant orbital interaction between them make this hexacoordinate assignment questionable.19 It was only very recently that a series of planar hexacoordinate carbon (phC) species, CE3M3+ (E = S–Te; M = Li–Cs), were designed computationally by the groups of Tiznado and Merino (Fig. 1; left panel),20 in which there exist pure electrostatic interactions between the negative Cδ− center and positive Mδ+ ligands. These phC clusters were achieved following the so-called “proper polarization of ligand” strategy.Open in a separate windowFig. 1The pictorial depiction of previously reported phC CE3M3+ (E = S–Te; M = Li–Cs) clusters and the present SiE3M3+ (E = S–Te and N–Sb; M = Li–Cs and Ca–Ba) clusters. Herein the solid and dashed lines represent covalent and ionic bonding, respectively. The opposite double arrows illustrate electrostatic repulsion.The concept of planar hypercoordinate carbons has been naturally extended to their next heavier congener, silicon-based systems. Although the steric repulsion between ligands decreases due to the larger size, the strength of π- and σ-bonding between the central atom and peripheral ligands dramatically decreases, which is crucial for stability. Planar tetracoordinate silicon (ptSi) was first experimentally observed in a pentaatomic C2v SiAl4− cluster by Wang and coworkers in 2000.21 Very recently, this topic got a huge boost by the room-temperature, large-scale syntheses of complexes containing a ptSi unit.22 A recent computational study also predicted the global minimum of SiMg4Y− (Y = In, Tl) and SiMg3In2 to have unprecendented planar pentacoordinate Si (ppSi) units.23 Planar hexacoordinate Si (phSi) systems seem to be even more difficult to stabilize. Previously, a C2v symmetric Cu6H6Si cluster was predicted as the true minimum,24 albeit its potential energy surface was not fully explored. A kinetically viable phSi SiAl3Mg3H2+ cluster cation was also predicted.25 However, these phSi systems24,25 are only local minima and not likely to be observed experimentally. In 2018, the group of Chen identified the Ca4Si22− building block containing a ppSi center and constructed an infinite CaSi monolayer, which is essentially a two-dimensional lattice of the Ca4Si2 motif.26 Thus, it is still an open question to achieve a phSi atom to date.Herein we have tried to find the correct combination towards a phSi system as the most stable isomer. Gratifyingly, we found a series of clusters, SiE3M3+ (E = N, P, As, Sb; M = Ca, Sr, Ba), having planar D3h symmetry with Si at the center of the six membered ring, as true global minimum forms. Si–E bonds are very strong in all the clusters, and alkaline-earth metals interact with the Si center by employing their d orbitals. However, electrostatic repulsion originated from the positively charged Si and M centers for E = N, P, and As dominates over attractive covalent interaction, making individual Si–M contacts repulsive in nature. This makes the assignment of SiE3M3+ (E = N, P, As; M = Ca, Sr, Ba) as genuine phSi somewhat skeptical. SiSb3M3+ (M = Ca, Sr, Ba) clusters are the sole candidates which possess genuine phSi centers as both electrostatic and covalent interactions in Si–M bonds are attractive. The d orbitals of M ligands play a crucial role in stabilizing the ligand framework and forming covalent bonds with phSi. Such planar hypercoordinate atoms are, in general, susceptible to external perturbations. However, the present title clusters maintain the planarity and the attractive nature of the bonds even after multiple ligand binding at M centers in SiSb3M3(NHC)6+ and SiSb3M3(Bz)6+. This would open the door for large-scale synthesis of phSi as well.Two major computational efforts were made before reaching our title phSi clusters. The first one is to examine SiE3M3+ (E = S–Po; M = Li–Cs) clusters, which adopt D3h or C3v structures as true minima (see Table S1 in ESI†), being isoelectronic to the previous phC CE3M3+ (E = S–Po; M = Li–Cs) clusters. In the SiE3M3+ (E = S–Po; M = Li–Cs) clusters, the Si center always carries a positive charge ranging from 0.01 to +1.03|e|, in contrast to the corresponding phC species (see Fig. 1). Thus, electrostatic interactions between the Siδ+ and Mδ+ centers would be repulsive (Fig. 1). Given that the possibility of covalent interaction with an alkali metal is minimal, it would be a matter of debate whether they could be called true coordination. A second effort is to tune the electronegativity difference between Si and M centers so that the covalent contribution in Si–M bonding becomes substantial. Along this line, we consider the combinations of SiE3M3+ (E = N, P, As, Sb; M = Be, Mg, Ca, Sr, Ba). The results in Fig. S1† show that for E = Be and Mg, the phSi geometry has a large out-of-plane imaginary frequency mode, which indicates a size mismatch between the Si center and peripheral E3M3 (E = N–Bi; M = Be, Mg) ring. On the other hand, the use of larger M = Ca, Sr, Ba atoms effectively expands the size of the cavity and eventually leads to perfect planar geometry with Si atoms at the center as minima. In the case of SiBi3M3+, the planar isomer possesses a small imaginary frequency for M = Ca. Although planar SiBi3Sr3+ and SiBi3Ba3+ are true minima, they are 2.2 and 2.5 kcal mol−1 higher in energy than the lowest energy isomer, respectively (Fig. S2†). Fig. 2 displays some selected low-lying isomers of SiE3M3+ (E = N, P, As, Sb; M = Ca, Sr, Ba) clusters (see Fig. S3–S6† for additional isomers). The global minimum structure is a D3h symmetric phSi with an 1A1′ electronic state for all the twelve cases. The second lowest energy isomer, a ppSi, is located more than 49 kcal mol−1 above phSi for E = N. This relative energy between the most stable and nearest energy isomer gradually decreases upon moving from N to Sb. In the case of SiSb3M3+ clusters, the second-lowest energy isomer is 4.6–6.1 kcal mol−1 higher in energy than phSi. The nearest triplet state isomer is very high in energy (by 36–53 kcal mol−1, Fig. S3–S6†) with respect to the global minimum.Open in a separate windowFig. 2The structures of low-lying isomers of SiE3M3+ (E = N, P, As, Sb; M = Ca, Sr, Ba) clusters. Relative energies (in kcal mol−1) are shown at the single-point CCSD(T)/def2-TZVP//PBE0/def2-TZVP level, followed by a zero-energy correction at PBE0. The values from left to right refer to Ca, Sr, and Ba in sequence. The group symmetries and electronic states are also given.Born–Oppenheimer molecular dynamics (BOMD) simulations at room temperature (298 K), taking SiE3Ca3+ clusters as case studies, were also performed. The results are displayed in Fig. S7.† All trajectories show no isomerization or other structural alterations during the simulation time, as indicated by the small root mean square deviation (RMSD) values. The BOMD data suggest that the global minimum also has reasonable kinetic stability against isomerization and decomposition.The bond distances, natural atomic charges, and bond indices for SiE3Ca3+ clusters are given in † for M = Sr, Ba). The Si–E bond distances are shorter than the typical Si–E single bond distance computed using the self-consistent covalent radii proposed by Pyykkö.27 In contrast, the Si–M bond distance is almost equal to the single bond distance. This gives the first hint of the presence of covalent bonding therein. However, the Wiberg bond indices (WBIs) for the Si–M links are surprisingly low (0.02–0.04). We then checked the Mayer bond order (MBO), which can be seen as a generalization of WBIs and is more acceptable since the approach of WBI calculations assumes orthonormal conditions of basis functions while the MBO considers an overlap matrix. The MBO values for the Si–M links are now sizable (0.13–0.18). These values are reasonable considering the large difference in electronegativity between Si and M, and, therefore, only a very polar bond is expected between them. In fact, the calculations of WBIs after orthogonalization of basis functions by the Löwdin method gives significantly large bond orders (0.48–0.55), which is known to overestimate the bond orders somewhat. The above results indicate that the presence of covalent bonding cannot be ruled out only by looking at WBI values.Bond distances (r, in Å), different bond orders (WBIs) {MBOs} [WBI in orthogonalized basis], and natural atomic charges (q, in |e|) of SiE3Ca3+ (E = N, P, As, Sb) clusters at the PBE0/def2-TZVP level
Open in a separate windowOur following argument regarding the presence of covalent Si–M bonding is based on energy decomposition analysis (EDA) in combination with natural orbital for chemical valence (NOCV) theory. We first performed EDA by taking Ca and SiE3Ca2 in different charge and electronic states as interacting fragments to get the optimum fragmentation scheme that suits the best to describe the bonding situation (see Tables S6–S9†). The size of orbital interaction (ΔEorb) is used as a probe.28 For all cases, Ca+ (D, 4s1) and SiE3Ca2 (D) in their doublet spin states turn out to be the best schemes, which give the lowest ΔEorb value. Energy term Interaction Ca+ (D, 4s1) + SiN3Ca2 (D) Ca+ (D, 4s1) + SiP3Ca2 (D) Ca+ (D, 4s1) + SiAs3Ca2 (D) Ca+ (D, 4s1) + SiSb3Ca2 (D) ΔEint −192.9 −153.0 −144.9 −129.9 ΔEPauli 139.8 115.2 115.7 110.9 ΔEelstata −162.0 (48.7%) −116.4 (43.4%) −113.0 (43.4%) −100.9 (41.9%) ΔEorba −170.7 (51.3%) −151.8 (56.6%) −147.6 (56.6%) −140.0 (58.1%) ΔEorb(1)b SiE3Ca2–Ca+(s) electron-sharing σ-bond −89.2 (52.3%) −79.4 (52.3%) −74.3 (50.3%) −66.9 (47.8%) ΔEorb(2)b SiE3Ca2 → Ca+(d) π‖-donation −32.9 (19.3%) −32.0 (21.1%) −31.8 (21.5%) −30.8 (22.0%) ΔEorb(3)b SiE3Ca2 → Ca+(d) σ-donation −13.1 (7.7%) −11.9 (7.8%) −12.0 (8.1%) −11.9 (8.5%) ΔEorb(4)b SiE3Ca2 → Ca+(d) π⊥-donation −12.3 (7.2%) −12.2 (8.0%) −12.5 (8.5%) −12.5 (8.9%) ΔEorb(5)b SiE3Ca2 → Ca+(d) δ-donation −8.1 (4.7%) −9.9 (6.5%) −10.9 (7.4%) −11.8 (8.4%) ΔEorb(rest)b −15.1 (8.8%) −6.4 (4.2%) −6.1 (4.1%) −6.1 (4.4%)
r Si–E | r Si–Ca | r E–Ca | q Si | q E | q Ca | |
---|---|---|---|---|---|---|
E = N | 1.669 | 2.555 | 2.246 | 1.57 | −1.93 | 1.74 |
(1.14) {1.23} [1.84] | (0.02) {0.13} [0.51] | (0.22) {0.67} [0.84] | ||||
E = P | 2.180 | 2.935 | 2.640 | 0.25 | −1.42 | 1.67 |
(1.34) {1.11} [1.52] | (0.03) {0.14} [0.54] | (0.27) {0.74} [1.05] | ||||
E = As | 2.301 | 3.004 | 2.721 | 0.07 | −1.34 | 1.65 |
(1.33) {1.10} [1.45] | (0.03) {0.15} [0.55] | (0.29) {0.71} [1.12] | ||||
E = Sb | 2.538 | 3.155 | 2.896 | −0.39 | −1.16 | 1.62 |
(1.29) {1.01} [1.33] | (0.04) {0.18} [0.48] | (0.30) {0.78} [1.14] |