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 828 毫秒
1.
Until now, all B≡B triple bonds have been achieved by adopting two ligands in the L→B≡B←L manner. Herein, we report an alternative route of designing the B≡B bonds based on the assumption that by acquiring two extra electrons, an element with the atomic number Z can have properties similar to those of the element with the atomic number Z+2. Specifically, we show that due to the electron donation from Al to B, the negatively charged B≡B kernel in the B2Al3 cluster mimics a triple N≡N bond. Comprehensive computational searches reveal that the global minimum structure of B2Al3 exhibits a direct B–B distance of 1.553 Å, and its calculated electron vertical detachment energies are in excellent agreement with the corresponding values of the experimental photoelectron spectrum. Chemical bonding analysis revealed one σ and two π bonds between the two B atoms, thus confirming a classical textbook B≡B triple bond, similar to that of N2.  相似文献   

2.
Despite its electron deficiency, boron is versatile in forming multiple bonds. Transition‐metal–boron double bonding is known, but boron–metal triple bonds have been elusive. Two bismuth boron cluster anions, BiB2O and Bi2B, containing triple and double B−Bi bonds are presented. The BiB2O and Bi2B clusters are produced by laser vaporization of a mixed B/Bi target and characterized by photoelectron spectroscopy and ab initio calculations. Well‐resolved photoelectron spectra are obtained and interpreted with the help of ab initio calculations, which show that both species are linear. Chemical bonding analyses reveal that Bi forms triple and double bonds with boron in BiB2O ([Bi≡B−B≡O]) and Bi2B ([Bi=B=Bi]), respectively. The Bi−B double and triple bond strengths are calculated to be 3.21 and 4.70 eV, respectively. This is the first experimental observation of Bi−B double and triple bonds, opening the door to design main‐group metal–boron complexes with multiple bonding.  相似文献   

3.
The asymmetric unit of the title compound, lead(II) dicalcium octaaluminate, contains one Pb, one Ca, four Al and eight O atoms, with the Pb atom and one O atom situated on mirror planes. Three Al atoms exhibit slightly distorted tetrahedral coordinations with a mean Al—O bond length of 1.76 Å. The fourth Al atom is in a considerably distorted trigonal–bipyramidal coordination with a mean Al—O bond length of 1.89 Å. One AlO4 tetrahedron forms infinite chains parallel to [100] via corner‐sharing. These chains are linked by parallel chains of edge‐sharing AlO5 trigonal bipyramids into layers A of six‐membered double rings extending parallel to (010). The second layer B is made up of the remaining two AlO4 tetrahedra. These tetrahedra share corners, resulting in likewise six‐membered double rings. Finally, the parallel layers A and B are linked into a three‐dimensional framework by common corners. Charge compensation is achieved by the Pb2+ and Ca2+ cations, which are situated in the cavities of the anionic framework, and which are surrounded by seven and six O atoms, respectively, both within highly irregular coordination polyhedra.  相似文献   

4.
The local angular distortions Δθ are theoretically studied for the various Ni3+ centers in LiAlyCo1–yO2 at different Al concentrations (y = 0, 0.1, 0.5, and 0.8) based on the perturbation calculations of electron paramagnetic resonance g factors for a trigonally distorted octahedral 3d7 cluster with low spin (S = 1/2). Due to the Jahn–Teller effect, the [NiO6]9– clusters are found to experience the local angular distortions (Δθ ≈ 5°–9°) along the C3 axis. The variation trend of Δθ with y is in accordance with that of anisotropy (Δg = g|| − g). As the substitutions can weaken bond strengths between transition metal and oxygen and the structural stability plays an important role in cathode performances, detailed investigations on the structural properties of the cathode materials LiAlyCo1–yO2 can be practically helpful to understand the performances of these materials. The oxy‐redox properties of LiAlyCo1–yO2 systems are comprehensible in the framework of Ni3+/Ni4+ couples, and the trigonally compressed octahedral [NiO6]9– clusters are applicable to the clarification of the electrochemical properties of lithium nickel oxide batteries. It appears that LiAl0.8Co0.2O2 with the largest Al concentration may correspond to the smallest distortion among the mixing systems.  相似文献   

5.
In the title compound, Na+·H+·2C8H7O3, the anion contains a short Speakman-type hydrogen bond [O⃛O = 2.413 (2) Å]. The anions and the Na atoms lie across twofold axes.  相似文献   

6.
Boragermene 3 featuring a double bond between the Ge and dicoordinate B atoms has been synthesized for the first time by reacting the cyclic (alkyl)(boryl)germylene–PMe3 adduct 1 with Cl2BN(SiMe3)2 followed by reductive dehalogenation with KC8. Addition of a Lewis base (MeNHC) to 3 leads to the formation of the corresponding adduct 4 , which shows double bond character between the Ge and tricoordinate B atoms. Compound 3 undergoes hydrogenation with H2 concomitant with a complete scission of the Ge=B bond.  相似文献   

7.
Boragermene 3 featuring a double bond between the Ge and dicoordinate B atoms has been synthesized for the first time by reacting the cyclic (alkyl)(boryl)germylene–PMe3 adduct 1 with Cl2BN(SiMe3)2 followed by reductive dehalogenation with KC8. Addition of a Lewis base (MeNHC) to 3 leads to the formation of the corresponding adduct 4 , which shows double bond character between the Ge and tricoordinate B atoms. Compound 3 undergoes hydrogenation with H2 concomitant with a complete scission of the Ge=B bond.  相似文献   

8.
Although nanometer-sized aluminum hydroxide clusters (i.e., ϵ-Al13, [Al13O4(OH)24(H2O)12]7+) command a central role in aluminum ion speciation and transformations between minerals, measurement of their translational diffusion is often limited to indirect methods. Here, 27Al pulsed field gradient stimulated echo nuclear magnetic resonance (PFGSTE NMR) spectroscopy has been applied to the AlO4 core of the ϵ-Al13 cluster with complementary theoretical simulations of the diffusion coefficient and corresponding hydrodynamic radii from a boundary element-based calculation. The tetrahedral AlO4 center of the ϵ-Al13 cluster is symmetric and exhibits only weak quadrupolar coupling, which results in favorable T1 and T2 27Al NMR relaxation coefficients for 27Al PFGSTE NMR studies. Stokes–Einstein relationship was used to relate the 27Al diffusion coefficient of the ϵ-Al13 cluster to the hydrodynamic radius for comparison with theoretical simulations, dynamic light scattering from literature, and previously published 1H PFGSTE NMR studies of chelated Keggin clusters. This first-of-its-kind observation proves that 27Al PFGSTE NMR diffusometry can probe symmetric Al environments in polynuclear clusters of greater molecular weight than previously considered.  相似文献   

9.
Hydrogen Bonds in o- and m-Phenylenediammonium Aquapentafluoro Metallates(III) (MIII = Al, Cr, Fe) m- and o-Phenylenediammonium-[MIIIF5(H2O)] compounds of Al, Cr and Fe were synthesized and characterized by X-ray single crystal structure analysis. All structures are described in the space group P212121 (Z = 4). m-Ph(NH3)22+ (Ph(NH3)22+ = phenylenediammonium) compounds: Al : a = 6.489(2), b = 7.943(2), c = 18.204(2) Å, R/wR = 0.084/0.050 for 1 533 reflections; Cr : a = 6.571(2), b = 8.006(2), c = 18.456(3) Å, R/wR = 0.050/0.040 for 1 571 reflections; Fe : a = 6.608(2), b = 8.052(2), c = 18.424(4) Å, R/wR = 0.042/0.034 for 1 947 reflections. o-Ph(NH3)22+ compounds: Al : a = 6.580(2), b = 7.891(2), c = 18.319(5) Å, R/wR = 0.050/0.045 for 2 370 reflections; Cr : a = 6.642(2), b = 7.954(2), c = 18.484(4) Å, R/wR = 0.065/0.043 for 2 041 reflections; Fe : a = 6.693(2), b = 7.995(4), c = 18.529(7) Å, R/wR = 0.035/0.033 for 2 651 reflections. Isolated distorted octahedral [MIIIF5(H2O)]2? anions are connected by double O? H ?F hydrogen bonds of alternating strength to form chains in the b direction. Those chains, packed in a pseudohexagonal way, are further linked by the ammonium functions of the phenylenediammonium cations to a 3 D hydrogen bond network.  相似文献   

10.
The asymmetric unit of the title compound, (C5H6N)2[NbCl4O(C5H5N)]Cl or (pyH)2[O=NbCl4(py)]Cl (py is pyridine), contains a discrete anionic niobium(V) complex, [O=NbCl4(py)], and two protonated pyridine mol­ecules, which form medium–strong hydrogen bonds with the Cl counter‐ion. The Nb=O distance of 1.7643 (17) Å is the longest among those in congener niobium complexes reported to date. Extensive density functional theory studies of conformations of [O=NbCl4(py)] and structural data mining raise doubts regarding the reliability of the length of this Nb=O double bond.  相似文献   

11.
《化学:亚洲杂志》2017,12(8):910-919
Reduction of aluminum(III), gallium(III), and indium(III) phthalocyanine chlorides by sodium fluorenone ketyl in the presence of tetrabutylammonium cations yielded crystalline salts of the type (Bu4N+)2[MIII(HFl−O)(Pc.3−)].−(Br) ⋅ 1.5 C6H4Cl2 [M=Al ( 1 ), Ga ( 2 ); HFl−O=fluoren‐9‐olato anion; Pc=phthalocyanine] and (Bu4N+) [InIIIBr(Pc.3−)].− ⋅ 0.875 C6H4Cl2 ⋅ 0.125 C6H14 ( 3 ). The salts were found to contain Pc.3− radical anions with negatively charged phthalocyanine macrocycles, as evidenced by the presence of intense bands of Pc.3− in the near‐IR region and a noticeable blueshift in both the Q and Soret bands of phthalocyanine. The metal(III) atoms coordinate HFl−O anions in 1 and 2 with short Al−O and Ga−O bond lengths of 1.749(2) and 1.836(6) Å, respectively. The C−O bonds [1.402(3) and 1.391(11) Å in 1 and 2 , respectively] in the HFl−O anions are longer than the same bond in the fluorenone ketyl (1.27–1.31 Å). Salts 1 – 3 show effective magnetic moments of 1.72, 1.66, and 1.79 μB at 300 K, respectively, owing to the presence of unpaired S= 1/2 spins on Pc.3−. These spins are coupled antiferromagnetically with Weiss temperatures of −22, −14, and −30 K for 1 – 3 , respectively. Coupling can occur in the corrugated two‐dimensional phthalocyanine layers of 1 and 2 with an exchange interaction of J /k B=−0.9 and −1.1 K, respectively, and in the π‐stacking {[InIIIBr(Pc.3−)].−}2 dimers of 3 with an exchange interaction of J /k B=−10.8 K. The salts show intense electron paramagnetic resonance (EPR) signals attributed to Pc.3−. It was found that increasing the size of the central metal atom strongly broadened these EPR signals.  相似文献   

12.
The effect of the anion charge on the structure of [LiAl2(OH)6]nX (X = Cl-, Br-, I-, SO4 2-, C6H8O4 2-) intercalation compounds and the water effect on the structure of [LiAl2(OH)6]Cl·nH2O have been studied using 1H, 7Li, and 27Al NMR. A change in the charge on the anion leads to significant changes in the asymmetry parameter for the lithium and aluminum nuclei with relatively small changes in the quadrupolar coupling constant and the broadening factor. The structure of the intermediate [LiAl2(OH)6]Cl·0.5H2O hydrate can be represented as a derivative of the structure of the anhydrous [LiAl2(OH)6]Cl intercalate with a slightly increased layer thickness and a minor orthorhombic distortion of the hexagonal cell; the water molecules partially fill the interlayer voids and participate in the diffusion process. Further hydration of the intercalate (x 1) leads to a minor (0.2) increase in the layer thickness and is accompanied by disordering of chloride ions and water molecules in the interlayer space.  相似文献   

13.
Schnöckel's [(AlCp*)4] and Jutzi's [SiCp*][B(C6F5)4] (Cp*=C5Me5) are landmarks in modern main-group chemistry with diverse applications in synthesis and catalysis. Despite the isoelectronic relationship between the AlCp* and the [SiCp*]+ fragments, their mutual reactivity is hitherto unknown. Here, we report on their reaction giving the complex salts [Cp*Si(AlCp*)3][WCA] ([WCA]=[Al(ORF)4] and [F{Al(ORF)3}2]; RF=C(CF3)3). The tetrahedral [SiAl3]+ core not only represents a rare example of a low-valent silicon-doped aluminium-cluster, but also—due to its facile accessibility and high stability—provides a convenient preparative entry towards low-valent Si−Al clusters in general. For example, an elusive binuclear [Si2(AlCp*)5]2+ with extremely short Al−Si bonds and a high negative partial charge at the Si atoms was structurally characterised and its bonding situation analysed by DFT. Crystals of the isostructural [Ge2(AlCp*)5]2+ dication were also obtained and represent the first mixed Al−Ge cluster.  相似文献   

14.
Single-atom catalysts are promising platforms for heterogeneous catalysis, especially for clean energy conversion, storage, and utilization. Although great efforts have been made to examine the bonding and oxidation state of single-atom catalysts before and/or after catalytic reactions, when information about dynamic evolution is not sufficient, the underlying mechanisms are often overlooked. Herein, we report the direct observation of the charge transfer and bond evolution of a single-atom Pt/C3N4 catalyst in photocatalytic water splitting by synchronous illumination X-ray photoelectron spectroscopy. Specifically, under light excitation, we observed Pt−N bond cleavage to form a Pt0 species and the corresponding C=N bond reconstruction; these features could not be detected on the metallic platinum-decorated C3N4 catalyst. As expected, H2 production activity (14.7 mmol h−1 g−1) was enhanced significantly with the single-atom Pt/C3N4 catalyst as compared to metallic Pt-C3N4 (0.74 mmol h−1 g−1).  相似文献   

15.
The tetracationic, univalent cluster compounds [{M(dmpe)}4]4+ (M=Ga, In; dmpe=bis(dimethylphosphino)ethane) were synthesized as their pf salts ([pf]=[Al(ORF)4]; RF=C(CF3)3). The four-membered ring in [{M(dmpe)}4]4+ is slightly puckered for M=Ga and almost square planar for M=In. Yet, although structurally similar, only the gallium cluster is prevalent in solution, while the indium cluster forms temperature dependent equilibria that include even the monomeric cation [In(dmpe)]+. This system is the first report of one and the same ligand inducing formation of isoelectronic and isostructural gallium/indium cluster cations. The system allows to study systematically analogies and differences with thermodynamic considerations and bonding analyses, but also to outline perspectives for bond activation using cationic, subvalent group 13 clusters.  相似文献   

16.
The chemistry of the low‐valent Group 13 elements (E = B, Al, Ga, In, Tl) has formed the recent hot topic. Recently, a series of low‐valent Group 13‐based compounds have been synthesized, i.e., [E‐Cp*‐E]+ (E = Al, Ga, In, Tl) cations, which have been termed as the interesting “inverse sandwich” complexes. To enrich the family of inverse sandwiches, we report our theoretical design of a new type of inverse sandwiches E‐C4H4‐E (E = Al, Ga, In, Tl) for stabilizing the low‐valent Group 13 elements. The calculated dissociation energies indicate that unlike [E‐Cp‐E]+ that dissociates via loss of the charged atom E+, E‐C4H4‐E dissociates via loss of the neutral atom E with the bond strengths of Al > Ga > In > Tl. Moreover, E‐C4H4‐E are more stable in dissociation than [E‐Cp‐E]+ cations. By comparing with other various isomers, we found that the inverted E‐C4H4‐E should be kinetically quite stable with the least conversion barriers of 33.5, 33.5, 35.2, and 36.9 kcal/mol for E = Al, Ga, In, and Tl, respectively. Furthermore, replacement of cyclobutadiene‐H atoms by the highly electron‐positive groups such as SiH3 and Si(CH3)3 could significantly stabilize the inverted form in thermodynamics. Possible synthetic routes are proposed for E‐C4H4‐E. With no need of counterions, the newly designed neutral complexes E‐C4H4‐E welcome future synthesis. © 2012 Wiley Periodicals, Inc.  相似文献   

17.
Geometry optimization of small (H2O)n+ clusters (n ≤ 4) at the UHF/4–31 + + G** level indicates that the cations consist of two fragments: the OH radical and the H2n−1 O+n−1 ion. The latter can be considered as a thermodynamically stable combination of a distorted H3O+ ion and (n−2) H2O molecules. The H bond between the fragments becomes weaker with increasing cluster size. Extrapolation of the adiabatic ionization potentials calculated for the (H2O)n oligomers (n ≤ 4) at the MP2 level to an infinite cluster size provides the value of approximately 8.7 eV, which can be presumably necessary for the ionization of liquid water in a vacuum. © 1997 John Wiley & Sons, Inc.  相似文献   

18.
Ternary clusters (NH3)·(H2SO4)·(H2O)n have been widely studied. However, the structures and binding energies of relatively larger cluster (n > 6) remain unclear, which hinders the study of other interesting properties. Ternary clusters of (NH3)·(H2SO4)·(H2O)n, n = 0-14, were investigated using MD simulations and quantum chemical calculations. For n = 1, a proton was transferred from H2SO4 to NH3. For n = 10, both protons of H2SO4 were transferred to NH3 and H2O, respectively. The NH4+ and HSO4 formed a contact ion-pair [NH4+-HSO4] for n = 1-6 and a solvent separated ion-pair [NH4+-H2O-HSO4] for n = 7-9. Therefore, we observed two obvious transitions from neutral to single protonation (from H2SO4 to NH3) to double protonation (from H2SO4 to NH3 and H2O) with increasing n. In general, the structures with single protonation and solvated ion-pair were higher in entropy than those with double protonation and contact ion-pair of single protonation and were thus preferred at higher temperature. As a result, the inversion between single and double protonated clusters was postponed until n = 12 according to the average binding Gibbs free energy at the normal condition. These results can serve as a good start point for studies of the other properties of these clusters and as a model for the solvation of the [H2SO4-NH3] complex in bulk water.  相似文献   

19.
The Crystal Structure of (Al0,5Ga0,5)CuOAsO4 – Copper Intermediate between Planar and Closed Coordination Single crystals of the new oxide arsenate (Al0.5Ga0.5)CuOAsO4 (monoclinic, P21/c, a = 734.3(2) pm, b = 1024.79(9) pm, c = 563.4(2) pm, β = 99.93(1)°, Z = 4) were obtained by reaction of Al/As/Cu/Ga-alloys with oxygen. The crystal structure was determined from four-circle diffractometer data (w2R = 0.039 for 1211 F2 values and 76 parameters). The structure contains [Cu2O6] double squares arranged in slabs perpendicular to the a axis such that a [4 + 1]-coordination of the copper atoms by oxygen atoms results which is intermediate between square-planar and square-pyramidal. Along [100] layers of corner sharing (Al/Ga)O4 and AsO4 tetrahedra are alternating with buckled Cu layers.  相似文献   

20.
The forced hydrolysis reaction of aqueous aluminum ion (Al3+) is of critical importance in Al chemistry, but its microscopic mechanism has long been neglected. Herein, density functional calculations reveal an external OH‐induced barrierless proton dissociation mechanism for the forced hydrolysis of Al3+(aq). Dynamic reaction pathway modeling results show that the barrierless deprotonations induced by the second‐ or third‐shell external OH proceed via the concerted proton transfer through H‐bond wires connected to the coordinated waters, and the inducing ability of the external OH decreases with increasing hydration layers between Al(H2O)63+ and the external OH. The OH‐induced forced hydrolysis mechanism of Al3+(aq) is quite different from its self‐hydrolysis mechanism without OH. The inducing ability is a unique characteristic of OH, rather than other anions such as F or Cl.  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