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 437 毫秒
1.
To gain more insight into the reactivity of intermetalloid clusters, the reactivity of the Zintl phase K12Sn17, which contains [Sn4]4? and [Sn9]4? cluster anions, was investigated. The reaction of K12Sn17 with gold(I) phosphine chloride yielded K7[(η2‐Sn4)Au(η2‐Sn4)](NH3)16 ( 1 ) and K17[(η2‐Sn4)Au(η2‐Sn4)]2(NH2)3(NH3)52 ( 2 ), which both contain the anion [(Sn4)Au(Sn4)]7? ( 1 a ) that consists of two [Sn4]4? tetrahedra linked through a central gold atom. Anion 1 a represents the first binary Au?Sn polyanion. From this reaction, the solvate structure [K([2.2.2]crypt)]3K[Sn9](NH3)18 ( 3 ; [2.2.2]crypt=4,7,13,16,21,24‐hexaoxa‐1,10‐diazabicyclo[8.8.8]hexacosane) was also obtained. In the analogous reaction of mesitylcopper with K12Sn17 in the presence of [18]crown‐6 in liquid ammonia, crystals of the composition [K([18]crown‐6)]2[K([18]crown‐6)(MesH)(NH3)][Cu@Sn9](thf) ( 4 ) were isolated ([18]crown‐6=1,4,7,10,13,16‐hexaoxacyclooctadiene, MesH=mesitylene, thf=tetrahydrofuran) and featured a [Cu@Sn9]3? cluster. A similar reaction with [2.2.2]crypt as a sequestering agent led to the formation of crystals of [K[2.2.2]crypt][MesCuMes] ( 5 ). The cocrystallization of mesitylene in 4 and the presence of [MesCuMes]? ( 5 a ) in 5 provides strong evidence that the migration of a bare Cu atom into an Sn9 anion takes place through the release of a Mes? anion from mesitylcopper, which either migrates to another mesitylcopper to form 5 a or is subsequently protonated to give MesH.  相似文献   

2.
A new phase in europium‐tin‐chalcogenide chemistry has been prepared using the reactive flux method: Eu8(Sn4Se14)(Se3)2. The compound crystallizes in the orthorhombic space group P21212 with cell parameters a = 11.990(2) Å, b = 16.425(4) Å, c = 8.543(1) Å, and Z = 2. Eu8(Sn4Se14)(Se3)2 is a three dimensional structure with EuII cations linked together with an unusual (Sn4Se14)12– anionic unit and (Se3)2– chains. UV‐VIS‐NIR band‐gap analysis shows that these black metallic crystals are likely semiconductors with an optical band‐gap of 1.07 eV.  相似文献   

3.
The reactivity of TiCp2Cl2 (d0) towards Zintl clusters was studied in liquid ammonia (Cp=cyclopentadienyl). Reduction of TiIVCp2Cl2 and ligand exchange led to the formation of [TiIIICp2(NH3)2]+, also obtainable by recrystallization of [CpTiIIICl]2. Upon reaction with [K4Sn9], ligand exchange leads to [TiCp21‐Sn9)(NH3)]3?. A small variation of the stoichiometry led to the formation of [Ti(η4‐Sn8)Cp]3?, which cocrystallizes with [TiCp2(NH3)2]+ and [TiCp21‐Sn9)(NH3)]3?. Finally, the large intermetalloid cluster anion [Ti4Sn15Cp5]n? (n=4 or 5) was obtained from the reaction of K12Sn17 and TiCp2Cl2 in liquid ammonia. The isolation of three side products, [K([18]crown‐6)]Cp, [K([18]crown‐6)]Cp(NH3), and [K([2.2]crypt)]Cp, suggests a stepwise elimination of the Cl? and Cp? ligands from TiCp2Cl2 and thus gives a hint to the mechanism of the product formation in which [Ti(η4+2‐Sn8)Cp]3? has a key role.  相似文献   

4.
First‐principles calculations of the atomic and electronic structure of double‐wall nanotubes (DWNTs) of α‐V2O5 are performed. Relaxation of the DWNT structure leads to the formation of two types of local regions: 1) bulk‐type regions and 2) puckering regions. Calculated total density of states (DOS) of DWNTs considerably differ from that of single‐wall nanotubes and the single layer, as well as from the DOS of the bulk and double layer. Small shoulders that appear on edges of valence and conduction bands result in a considerable decrease in the band gaps of the DWNTs (up to 1 eV relative to the single‐layer gaps). The main reason for this effect is the shift of the inner‐ and outer‐wall DOS in opposite directions on the energetic scale. The electron density corresponding to shoulders at the conduction‐band edges is localized on vanadium atoms of the bulk‐type regions, whereas the electron density corresponding to shoulders at the valence‐band edges belongs to oxygen atoms of both regions.  相似文献   

5.
[AuII([12]anS4)]2+ – X‐ and Q‐Band EPR Evidence of a New Monomeric Gold(II) Compound The reaction of [AuIIICl4] with the thiacrown ether [12]aneS4 leads to an instable [AuII([12]anS4)]2+ complex (5d9, S = 1/2) which was characterized by X‐ and Q‐ band EPR. The spin Hamiltonian parameters g , A Au and P Au were derived using a program package allowing an exact diagonalisation of the spin‐Hamiltonian‐Matrix. The EPR parameters suggest the coordination of only one thiacrown ether ligand in the new AuII complex.  相似文献   

6.
The electrical conductivities and plausible charge‐ordering states in the room temperature (r.t.) phase for MMX chains [Ni2(dta)4I] and [Pt2(dta)4I] (dta = CH3CS) have been analyzed with periodic density functional theory (DFT) and correlated ab initio calculations combined with the effective Hamiltonian theory. Periodic DFT calculations show a more delocalized nature of the ground state in [Pt2(dta)4I] compared to [Ni2(dta)4I], which features a rather large energy gap between the occupied and empty bands, and charge polarized dimer units. A larger electrical conductivity for the Pt chain can be expected, especially because the Fermi level lies within a band with contributions from Pt and I orbitals. Electronic structure parameters extracted from ab initio cluster calculations show that the large difference between the observed conductivities at 300 K for Ni and Pt compounds, of 3 orders of magnitude, cannot be explained from the parameters extracted from an embedded M2(dta)4I2 dimer fragment alone. When tetramer fragments are considered, we observe that the interdimer transfer integral (t) between neighboring M2 units connected by an iodine atom at correlated level is comparable in both chains. On the other hand, the energy to transfer an electron from a dimer to the neighboring one (Coulomb repulsion U) is three times larger in the Ni compound with respect to the Pt chain, in line with the poor conductivity of the former. The electronic structure of the M4(dta)8I3 fragment points to an alternate charge‐polarization state for Ni and an average valence state for Pt when the r.t. X‐ray structure is considered. © 2012 Wiley Periodicals, Inc.  相似文献   

7.
In all known Group 5 transition‐metal dichalcogenide monolayers (MLs), the metal centers carry a spin, and their ground‐state phases are either metallic or semiconducting with indirect band gaps. Here, on grounds of first‐principles calculations, we report that the Haeckelite polytypes 1 S ‐NbX2 (X=S, Se, Te) are diamagnetic direct‐band‐gap semiconductors even though the Nb atoms are in the 4+ oxidation state. In contrast, 1 S ‐VX2 MLs are antiferromagnetically coupled indirect‐band‐gap semiconductors. The 1 S phases are thermodynamically and dynamically stable but of slightly higher energy than their 1 H and 1 T ML counterparts. 1 S ‐NbX2 MLs are excellent candidates for optoelectronic applications owing to their small band gaps (between 0.5 and 1 eV). Moreover, 1 S ‐NbS2 shows a particularly high hole mobility of 2.68×103 cm2 V−1 s−1, which is significantly higher than that of MoS2 and comparable to that of WSe2.  相似文献   

8.
The typical two‐dimensional (2D) semiconductors MoS2, MoSe2, WS2, WSe2 and black phosphorus have garnered tremendous interest for their unique electronic, optical, and chemical properties. However, all 2D semiconductors reported thus far feature band gaps that are smaller than 2.0 eV, which has greatly restricted their applications, especially in optoelectronic devices with photoresponse in the blue and UV range. Novel 2D mono‐elemental semiconductors, namely monolayered arsenene and antimonene, with wide band gaps and high stability were now developed based on first‐principles calculations. Interestingly, although As and Sb are typically semimetals in the bulk, they are transformed into indirect semiconductors with band gaps of 2.49 and 2.28 eV when thinned to one atomic layer. Significantly, under small biaxial strain, these materials were transformed from indirect into direct band‐gap semiconductors. Such dramatic changes in the electronic structure could pave the way for transistors with high on/off ratios, optoelectronic devices working under blue or UV light, and mechanical sensors based on new 2D crystals.  相似文献   

9.
In this work, a pincer‐type complex [Cp*Ir‐(SNPh)(SNHPh)(C2B10H9)] ( 2 ) was synthesized and its reactivity studied in detail. Interestingly, molecular hydrogen can induce the transformation between the metalloradical [Cp*Ir‐(SNPh)2(C2B10H9)] ( 5 .) and 2 . A mixed‐valence complex, [(Cp*Ir)2‐(SNPh)2(C2B10H8)] ( 7 .+), was also synthesized by one‐electron oxidation. Studies show that 7 .+ is fully delocalized, possessing a four‐centered‐one‐electron (S‐Ir‐Ir‐S) bonding interaction. DFT calculations were also in good agreement with the experimental results.  相似文献   

10.
In the title compound, [Cu(CN)(C4H5N3)]n or [Cu(μ‐CN)(μ‐PyzNH2)]n (PyzNH2 is 2‐aminopyrazine), the CuI center is tetrahedrally coordinated by two cyanide and two PyzNH2 ligands. The CuI–cyano links give rise to [Cu–CN] chains running along the c axis, which are bridged by bidentate PyzNH2 ligands. The three‐dimensional framework can be described as being formed by two interpenetrated three‐dimensional honeycomb‐like networks, both made of 26‐membered rings of composition [Cu6(μ‐CN)2(μ‐PyzNH2)4].  相似文献   

11.
A novel tetraoxolene‐bridged Fe two‐dimensional honeycomb layered compound, (NPr4)2[Fe2(Cl2An)3] ?2 (acetone)?H2O ( 1 ), where Cl2Ann?=2,5‐dichloro‐3,6‐dihydroxy‐1,4‐benzoquinonate and NPr4+=tetrapropylammonium cation, has been synthesized. 1 revealed a thermally induced valence tautomeric transition at T1/2=236 K (cooling)/237 K (heating) between Fem+ (m=2 or 3) and Cl2Ann? (n=2 or 3) that induced valence modulations between [FeIIHSFeIIIHS(Cl2An2?)2(Cl2An.3?)]2? at T>T1/2 and [FeIIIHSFeIIIHS(Cl2An2?)(Cl2An.3?)2]2? at T<T1/2. Even in a two‐dimensional network structure, the low‐temperature phase [FeIIIHSFeIIIHS(Cl2An2?)(Cl2An.3?)2]2? valence set can be regarded as a magnetic chain‐knit network, where ferrimagnetic Δ and Λ chains of [FeIIIHS(Cl2An.3?)] are alternately linked by the diamagnetic Cl2An2?. This results in a slow magnetization behavior attributed to the structure acting as a single‐chain magnet at lower temperatures.  相似文献   

12.
The compounds poly[di‐μ4‐succinato‐μ2‐1,2‐di‐4‐pyridylethane‐dicopper(II)], [Cu2(C4H4O4)2(C12H12N2)]n, (I), and poly[di‐μ4‐succinato‐μ2‐1,3‐di‐4‐pyridylpropane‐dicopper(II)], [Cu2(C4H4O4)2(C13H14N2)]n, (II), exhibit polymeric structures with the dicopper units doubly bridged by bis‐bidentate succinate groups and crosslinked by the separator bis(pyridyl) molecules. In (I), the molecule exhibits a centre of inversion located midway between the core Cu‐dimer atoms and another that relates half of the bis(pyridyl)ethane ligand to the other half. Compound (II) has a similar molecular packing but with a doubled lattice constant and noncentrosymmetric core units. An antiferromagnetic interaction due to the dinuclear copper units was deduced from magnetic subsceptibility measurements, and spin triplet signals were detected in the electron paramagnetic resonance spectra for both compounds.  相似文献   

13.
A facile synthesis of the [ReF6]2? ion and its use as a building block to synthesize magnetic systems are reported. Using dc and ac magnetic susceptibility measurements, INS and EPR spectroscopies, the magnetic properties of the isolated [ReF6]2? unit in (PPh4)2[ReF6]?2 H2O ( 1 ) have been fully studied including the slow relaxation of the magnetization observed below ca. 4 K. This slow dynamic is preserved for the one‐dimensional coordination polymer [Zn(viz)4(ReF6)] ( 2 , viz=1‐vinylimidazole), demonstrating the irrelevance of low symmetry for such magnetization dynamics in systems with easy‐plane‐type anisotropy. The ability of fluoride to mediate significant exchange interactions is exemplified by the isostructural [Ni(viz)4(ReF6)] ( 3 ) analogue in which the ferromagnetic NiII–ReIV interaction (+10.8 cm?1) dwarfs the coupling present in related cyanide‐bridged systems. These results reveal [ReF6]2? to be an unique new module for the design of molecule‐based magnetic materials.  相似文献   

14.
The Pb‐V oxyhalide apatite compounds Pb5(VO4)3X (X=F, Cl, Br, I) were successfully synthesized using a facile solution method and studied with respect to their structural/optical characteristics and electronic band structures. UV‐visible diffuse reflectance spectroscopy, electrochemical analysis and first‐principles calculations showed that the synthesized apatites behaved as n‐type semiconductors, with absorption bands in the UV‐visible region that could be assigned to electron transitions from the valence band to a conduction band formed by hybridized V 3d and Pb 6p orbitals. Among the apatites examined, Pb5(VO4)3I had the smallest band gap of 2.7 eV, due to an obvious contribution of I 5p orbitals to the valence band maximum. Based on its visible light absorption capability, Pb5(VO4)3I generated a continuous anodic photocurrent under visible light (λ>420 nm) in a solution of 0.1 m NaI in acetonitrile.  相似文献   

15.
The electronic structure of zeolite A is developed in a step by step procedure from the simple OhH8Si8O12 molecule, to the 1 [(-O)2H4[Si8O12)] chain, to the 2 [(-O4)(Si8O12)] layer, and finally to the silica zeolite A framework 3 [Si24O48]. It is remarkable how well the calculated band structures of both, 2 [(-O)4(Si8O12)] and 3 [Si24O48] correspond to the experimentally determined band structure of α-quartz with a Fermi level of -10.55 eV. The HOMO region consists in each case of nonbonding 2p-oxygen bands which in a localized language can be denoted as oxygen lone pairs ( | O<). We observe in each case the typical behaviour of an insulator with saturated valencies whose electronic structure can be described as being localized and is already present in the starting Oh-H8Si8O12 molecule. The double-8-rings D8R of the 2 [(-O)4(Si8O12)] layer have a pore diameter of 4.1 Å, the same as the pore opening of zeolite A. It is large enough to accept up to four Ag, forming 2 [(-O)4(Si8O12)Agn], n = 1, 2, 3, 4, layers, suitable for modelling the electronic interactions between the zeolite cavity embedded silver clusters and between the clusters and the zeolite framework. With one Ag per D8R the band structure is simply a superposition of the 4d, 5s and 5p levels of a layer of nearly noninteracting Ag and the silicon dioxide layer. The Ag-d band lies below the oxygen lone pairs, the Ag-s band lies about 3 eV above the oxygen lone pairs, and the Ag-5p bands are in the antibonding silicon dioxide region. The first electronic transition is of oxygen lone pair to Ag-5s LMCT type. Increasing silver content results in progressive splitting of the 5σ Ag bands and shifts the first (Agm+ n)? ← (| O<) charge transfer transition to lower energies. The filled Ag 4d-bands lie always significantly below the (| O<) HOCOs (highest occupied crystal orbitals) but their band width increases with increasing silver content. In all cases the zeolite environment separates the Ag clusters through antibonding Ag-(← O<) interactions so that the coupling remains weak and it makes sense to describe the Ag clusters in the D8R as quantum dots weakly interacting with each other.  相似文献   

16.
New compounds [Ru(pap)2(L)](ClO4), [Ru(pap)(L)2], and [Ru(acac)2(L)] (pap=2‐phenylazopyridine, L?=9‐oxidophenalenone, acac?=2,4‐pentanedionate) have been prepared and studied regarding their electron‐transfer behavior, both experimentally and by using DFT calculations. [Ru(pap)2(L)](ClO4) and [Ru(acac)2(L)] were characterized by crystal‐structure analysis. Spectroelectrochemistry (EPR, UV/Vis/NIR), in conjunction with cyclic voltammetry, showed a wide range of about 2 V for the potential of the RuIII/II couple, which was in agreement with the very different characteristics of the strongly π‐accepting pap ligand and the σ‐donating acac? ligand. At the rather high potential of +1.35 V versus SCE, the oxidation of L? into L. could be deduced from the near‐IR absorption of [RuIII(pap)(L.)(L?)]2+. Other intense long‐wavelength transitions, including LMCT (L?→RuIII) and LL/CT (pap.?→L?) processes, were confirmed by TD‐DFT results. DFT calculations and EPR data for the paramagnetic intermediates allowed us to assess the spin densities, which revealed two cases with considerable contributions from L‐radical‐involving forms, that is, [RuIII(pap0)2(L?)]2+?[RuII(pap0)2(L.)]2+ and [RuIII(pap0)(L?)2]+?[RuII(pap0)(L?)(L?)]+. Calculations of electrogenerated complex [RuII(pap.?)(pap0)(L?)] displayed considerable negative spin density (?0.188) at the bridging metal.  相似文献   

17.
The crystal structure of the low‐spin (S = 1) MnIII complex [Mn(CN)2(C10H24N4)]ClO4, or trans‐[Mn(CN)2(cyclam)](ClO4) (cyclam is the tetradentate amine ligand 1,4,8,11‐tetra­aza­cyclo­tetra­decane), is reported. The structural parameters in the Mn(cyclam) moiety are found to be insensitive to both the spin and the oxidation state of the Mn ion. The difference between high‐ and low‐spin MnIII complexes is that a pronounced tetragonal elongation of the coordination octahedron occurs in high‐spin complexes and a slight tetragonal compression is seen in low‐spin complexes, as in the title complex.  相似文献   

18.
Mercury(II) Chloride and Iodide Complexes of Dithia‐ and Tetrathiacrown Ethers The complexes [(HgCl2)2((ch)230S4O6)] ( 1 ), [HgCl2(mn21S2O5)] ( 2 ), [HgCl2(ch18S2O4)] ( 3 ) and [HgI(meb12S2O2)]2[Hg2I6] ( 4 ) have been synthesized, characterized and their crystal structures were determined. In [(HgCl2)2((ch)230S4O6)] two HgCl2 units are discretely bonded within the ligand cavity of the 30‐membered dichinoxaline‐tetrathia‐30‐crown‐10 ((ch)230S4O6) forming a binuclear complex. HgCl2 forms 1 : 1 “in‐cavity” complexes with the 21‐membered maleonitrile‐dithia‐21‐crown‐7 (mn21S2O5) ligand and the 18‐membered chinoxaline‐dithia‐18‐crown‐6 (ch18S2O4) ligand, respectively. The 12‐membered 4‐methyl‐benzo‐dithia‐12‐crown‐4 (meb12S2O2) ligand gave with two equivalents HgI2 the compound [HgI(meb12S2O2)]2[Hg2I6]. In the cation [HgI(meb12S2O2)]+ meb12S2O2 forms with the cation HgI+ a half‐sandwich complex.  相似文献   

19.
The title pendent‐arm macrocyclic hexa­amine ligand binds stereospecifically in a hexadentate manner, and we report here its isomorphous NiII and ZnII complexes (both as perchlorate salts), namely (cis‐6,13‐di­methyl‐1,4,8,11‐tetra­aza­cyclo­tetra­decane‐6,13‐di­amine‐κ6N)­nickel(II) di­per­chlorate, [Ni(C12H30N6)]­­(ClO4)2, and (cis‐6,13‐di­methyl‐1,4,8,11‐tetraaza‐cyclo­tetra­decane‐6,13‐di­amine‐κ6N)­zinc(II) di­per­chlorate, [Zn(C12H30N6)]­(ClO4)2. Distortion of the N—M—N valence angles from their ideal octahedral values becomes more pronounced with increasing metal‐ion size and the present results are compared with other structures of this ligand.  相似文献   

20.
Synthesis and Crystal Structure of the Mixed Valent Complex [Sn2I3(NPPh3)3] The mixed valent phosphoraneiminato complex [Sn2I3(NPPh3)3] ( 1 ) was prepared by the reaction of the tin(II) complex [SnI(NPPh3)]2 with sodium in tetrahydrofuran. 1 crystallizes with two formula units of THF to form yellow, moisture sensitive single crystals, which were characterized by a crystal structure determination. 1 · 2 THF: Space group P21/c, Z = 4, lattice dimensions at –80 °C: a = 1964.5(2), b = 1766.0(2), c = 2058.6(2) pm; β = 118.33(1)°, R = 0.052. 1 forms dimeric molecules in which the tin atoms are linked by two nitrogen atoms of two (NPPh3) groups to form a planar Sn2N2 four‐membered ring. The SnIV atom is additionally coordinated by a terminal iodine atom and by a terminal (NPPh3) group, whereas the SnII atom is additionally coordinated by two iodine atoms forming a ψ trigonal‐bipyramidal surrounding.  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